19.01.2013 Views

OXIDATIVE STRESS AND DISEASE.pdf - E-Lib FK UWKS

OXIDATIVE STRESS AND DISEASE.pdf - E-Lib FK UWKS

OXIDATIVE STRESS AND DISEASE.pdf - E-Lib FK UWKS

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>OXIDATIVE</strong> <strong>STRESS</strong> <strong>AND</strong> <strong>DISEASE</strong><br />

Series Editors<br />

LESTER PACKER, PH.D.<br />

ENRIQUE CADENAS, M.D., PH.D.<br />

University of Southern California School of Pharmacy<br />

Los Angeles, California<br />

1. Oxidative Stress in Cancer, AIDS, and Neurodegenerative<br />

Diseases, edited by Luc Montagnier, René Olivier, and<br />

Catherine Pasquier<br />

2. Understanding the Process of Aging: The Roles of<br />

Mitochondria, Free Radicals, and Antioxidants, edited by<br />

Enrique Cadenas and Lester Packer<br />

3. Redox Regulation of Cell Signaling and Its Clinical Application,<br />

edited by Lester Packer and Junji Yodoi<br />

4. Antioxidants in Diabetes Management, edited by Lester Packer,<br />

Peter Rösen, Hans J. Tritschler, George L. King,<br />

and Angelo Azzi<br />

5. Free Radicals in Brain Pathophysiology, edited by<br />

Giuseppe Poli, Enrique Cadenas, and Lester Packer<br />

6. Nutraceuticals in Health and Disease Prevention, edited by<br />

Klaus Krämer, Peter-Paul Hoppe, and Lester Packer<br />

7. Environmental Stressors in Health and Disease, edited by<br />

Jürgen Fuchs and Lester Packer<br />

8. Handbook of Antioxidants: Second Edition, Revised<br />

and Expanded, edited by Enrique Cadenas and Lester Packer<br />

9. Flavonoids in Health and Disease: Second Edition, Revised<br />

and Expanded, edited by Catherine A. Rice-Evans<br />

and Lester Packer<br />

10. Redox–Genome Interactions in Health and Disease, edited by<br />

Jürgen Fuchs, Maurizio Podda, and Lester Packer<br />

11. Thiamine: Catalytic Mechanisms in Normal and Disease States,<br />

edited by Frank Jordan and Mulchand S. Patel<br />

12. Phytochemicals in Health and Disease, edited by Yongping Bao<br />

and Roger Fenwick<br />

13. Carotenoids in Health and Disease, edited by Norman I. Krinsky,<br />

Susan T. Mayne, and Helmut Sies


14. Herbal and Traditional Medicine: Molecular Aspects of Health,<br />

edited by Lester Packer, Choon Nam Ong, and Barry Halliwell<br />

15. Nutrients and Cell Signaling, edited by Janos Zempleni<br />

and Krishnamurti Dakshinamurti<br />

16. Mitochondria in Health and Disease, edited by<br />

Carolyn D. Berdanier<br />

17. Nutrigenomics, edited by Gerald Rimbach, Jürgen Fuchs,<br />

and Lester Packer<br />

18. Oxidative Stress, Inflammation, and Health, edited by<br />

Young-Joon Surh and Lester Packer<br />

19. Nitric Oxide, Cell Signaling, and Gene Expression, edited by<br />

Santiago Lamas and Enrique Cadenas<br />

20. Resveratrol in Health and Disease, edited by<br />

Bharat B. Aggarwal and Shishir Shishodia<br />

21. Oxidative Stress and Age-Related Neurodegeneration,<br />

edited by Yuan Luo and Lester Packer<br />

22. Molecular Interventions in Lifestyle-Related Diseases, edited by<br />

Midori Hiramatsu, Toshikazu Yoshikawa, and Lester Packer<br />

23. Oxidative Stress and Inflammatory Mechanisms in Obesity,<br />

Diabetes, and the Metabolic Syndrome, edited by<br />

Lester Packer and Helmut Sies


Oxidative Stress<br />

and Inflammatory<br />

Mechanisms in Obesity,<br />

Diabetes, and the<br />

Metabolic Syndrome<br />

edited by<br />

Lester Packer<br />

Helmut Sies<br />

Boca Raton London New York<br />

CRC Press is an imprint of the<br />

Taylor & Francis Group, an informa business<br />

v


CRC Press<br />

Taylor & Francis Group<br />

6000 Broken Sound Parkway NW, Suite 300<br />

Boca Raton, FL 33487-2742<br />

© 2008 by Taylor & Francis Group, LLC<br />

CRC Press is an imprint of Taylor & Francis Group, an Informa business<br />

No claim to original U.S. Government works<br />

Printed in the United States of America on acid-free paper<br />

10 9 8 7 6 5 4 3 2 1<br />

International Standard Book Number-10: 1-4200-4378-1 (Hardcover)<br />

International Standard Book Number-13: 978-1-4200-4378-5 (Hardcover)<br />

This book contains information obtained from authentic and highly regarded sources. Reprinted material<br />

is quoted with permission, and sources are indicated. A wide variety of references are listed. Reasonable<br />

efforts have been made to publish reliable data and information, but the author and the publisher cannot<br />

assume responsibility for the validity of all materials or for the consequences of their use.<br />

No part of this book may be reprinted, reproduced, transmitted, or utilized in any form by any electronic,<br />

mechanical, or other means, now known or hereafter invented, including photocopying, microfilming,<br />

and recording, or in any information storage or retrieval system, without written permission from the<br />

publishers.<br />

For permission to photocopy or use material electronically from this work, please access www.copyright.<br />

com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC) 222 Rosewood<br />

Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and<br />

registration for a variety of users. For organizations that have been granted a photocopy license by the<br />

CCC, a separate system of payment has been arranged.<br />

Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are<br />

used only for identification and explanation without intent to infringe.<br />

<strong>Lib</strong>rary of Congress Cataloging-in-Publication Data<br />

Oxidative stress and inflammatory mechanisms in obesity, diabetes, and the<br />

metabolic syndrome / edited by Lester Packer and Helmut Sies.<br />

p. ; cm. -- (Oxidative stress and disease ; 22)<br />

“A CRC title.”<br />

Includes bibliographical references and index.<br />

ISBN-13: 978-1-4200-4378-5 (hardcover : alk. paper)<br />

ISBN-10: 1-4200-4378-1 (hardcover : alk. paper)<br />

1. Oxidative stress. 2. Metabolic syndrome. 3. Obesity. 4. Diabetes. I. Packer,<br />

Lester. II. Sies, H. (Helmut), 1942- III. Series.<br />

[DNLM: 1. Metabolic Syndrome X--physiopathology. 2. Diabetes Mellitus, Type<br />

2--physiopathology. 3. Inflammation--physiopathology. 4. Obesity--physiopathology.<br />

5. Oxidative Stress--physiology. W1 OX626 v.22 2007 / WK 820 O98 2007]<br />

QP177.O935 2007<br />

616.3’9--dc22 2007005664<br />

Visit the Taylor & Francis Web site at<br />

http://www.taylorandfrancis.com<br />

and the CRC Press Web site at<br />

http://www.crcpress.com


Table of Contents<br />

Series Preface.......................................................................................................xi<br />

Preface.................................................................................................................xv<br />

About the Editors............................................................................................. xvii<br />

Contributors........................................................................................................xix<br />

SECTION I Oxidative Stress, Metabolic Syndrome,<br />

Obesity, Diabetes, and Uncoupling<br />

Proteins<br />

Chapter 1<br />

The Metabolic Syndrome Defined........................................................................3<br />

Neil J. Stone and Jennifer Berliner<br />

Chapter 2<br />

The Metabolic Syndrome: The Question of Balance between the<br />

Pro-Inflammatory Effect of Macronutrients and the Anti-Inflammatory<br />

Effect of Insulin ..................................................................................................15<br />

Paresh Dandona, Ajay Chaudhuri, Priya Mohanty, and Husam Ghanim<br />

Chapter 3<br />

The Role of Oxidative Stress in Diseases Associated with<br />

Overweight and Obesity .....................................................................................33<br />

Ginger L. Milne, Ling Gao, Joshua D. Brooks, and Jason D. Morrow<br />

Chapter 4<br />

Metabolic Syndrome Due to Early Life Nutritional Modifications...................47<br />

Malathi Srinivasan, Paul Mitrani, and Mulchand S. Patel<br />

Chapter 5<br />

Oxidative Stress and Antioxidants in the Perinatal Period ................................71<br />

Hiromichi Shoji, Yuichiro Yamashiro, and Berthold Koletzko<br />

vii


viii Oxidative Stress and Inflammatory Mechanisms<br />

Chapter 6<br />

Maternal Obesity, Glucose Intolerance, and Inflammation in Pregnancy .........93<br />

Janet C. King<br />

Chapter 7<br />

Obesity, Nutrigenomics, Metabolic Syndrome, and Type 2 Diabetes.............107<br />

David Heber<br />

Chapter 8<br />

Post-Prandial Endothelial Dysfunction, Oxidative Stress, and<br />

Inflammation in Type 2 Diabetes .....................................................................123<br />

Antonio Ceriello<br />

Chapter 9<br />

Obesity and Inflammation: Implications for Atherosclerosis ..........................139<br />

John Alan Farmer<br />

Chapter 10<br />

Oligomeric Composition of Adiponectin and Obesity.....................................167<br />

T. Bobbert and Joachim Spranger<br />

Chapter 11<br />

Insulin-Stimulated Reactive Oxygen Species and Insulin<br />

Signal Transduction...........................................................................................177<br />

Barry J. Goldstein, Kalyankar Mahadev, and Xiangdong Wu<br />

Chapter 12<br />

Intracellular Signaling Pathways and Peroxisome Proliferator-Activated<br />

Receptors in Vascular Health in Hypertension and in Diabetes ......................195<br />

Farhad Amiri, Karim Benkirane, and Ernesto L. Schiffrin<br />

Chapter 13<br />

Role of Uncoupling Protein 2 in Pancreatic β Cell Function:<br />

Secretion and Survival ......................................................................................211<br />

Jingyu Diao, Catherine B. Chan, and Michael B. Wheeler


Table of Contents ix<br />

SECTION II Influence of Dietary Factors,<br />

Micronutrients, and Metabolism<br />

Chapter 14<br />

Nutritional Modulation of Inflammation in Metabolic Syndrome...................227<br />

Uma Singh, Sridevi Devaraj, and Ishwarlal Jialal<br />

Chapter 15<br />

Dietary Fatty Acids and Metabolic Syndrome.................................................243<br />

Helen M. Roche<br />

Chapter 16<br />

Lipid-Induced Death of Macrophages: Implication for<br />

Destabilization of Atherosclerotic Plaques.......................................................251<br />

Oren Tirosh and Anna Aronis<br />

Chapter 17<br />

α-Lipoic Acid Prevents Diabetes Mellitus and Endothelial<br />

Dysfunction in Diabetes-Prone Obese Rats .....................................................261<br />

Woo Je Lee, Ki-Up Lee, and Joong-Yeol Park<br />

Chapter 18<br />

Lipoic Acid Blocks Obesity through Reduced Food Intake, Enhanced<br />

Energy Expenditure, and Inhibited Adipocyte Differentiation ........................273<br />

Jong-Min Park and An-Sik Chung<br />

Chapter 19<br />

Effects of Conjugated Linoleic Acid and Lipoic Acid on Insulin<br />

Action in Insulin-Resistant Obese Zucker Rats ...............................................289<br />

Erik J. Henriksen<br />

Chapter 20<br />

Trivalent Chromium Supplementation Inhibits Oxidative Stress,<br />

Protein Glycosylation, and Vascular Inflammation in High<br />

Glucose-Exposed Human Erythrocytes and Monocytes ..................................301<br />

Sushil K. Jain<br />

Index .................................................................................................................315


Series Preface<br />

OXYGEN BIOLOGY <strong>AND</strong> MEDICINE<br />

Through evolution, oxygen — itself a free radical — was chosen as the terminal<br />

electron acceptor for respiration. The two unpaired electrons of oxygen spin in<br />

the same direction; thus, oxygen is a biradical. Other oxygen-derived free radicals<br />

such as superoxide anion or hydroxyl radicals formed during metabolism or by<br />

ionizing radiation are stronger oxidants, i.e., endowed with higher chemical<br />

reactivities. Oxygen-derived free radicals are generated during metabolism and<br />

energy production in the body and are involved in regulation of signal transduction<br />

and gene expression, activation of receptors and nuclear transcription factors,<br />

oxidative damage to cell components, antimicrobial and cytotoxic actions of<br />

immune system cells, as well as in aging and age-related degenerative diseases.<br />

Conversely, cells conserve antioxidant mechanisms to counteract the effects of<br />

oxidants; these antioxidants may remove oxidants either in a highly specific<br />

manner (for example, by superoxide dismutases) or in a less specific manner (for<br />

example, through small molecules such as vitamin E, vitamin C, and glutathione).<br />

Oxidative stress as classically defined is an imbalance between oxidants and<br />

antioxidants. Overwhelming evidence indicates that oxidative stress can lead to<br />

cell and tissue injury. However, the same free radicals that are generated during<br />

oxidative stress are produced during normal metabolism and, as a corollary, are<br />

involved in both human health and disease.<br />

UNDERST<strong>AND</strong>ING <strong>OXIDATIVE</strong> <strong>STRESS</strong><br />

In recent years, the research disciplines interested in oxidative stress have grown<br />

and enormously increased our knowledge of the importance of the cell redox<br />

status and the recognition of oxidative stress as a process with implications for<br />

many pathophysiological states. From this multi- and inter-disciplinary interest<br />

in oxidative stress emerges a concept that attests to the vast consequences of<br />

the complex and dynamic interplay of oxidants and antioxidants in cellular and<br />

tissue settings. Consequently, our view of oxidative stress is growing in scope<br />

and new future directions. Likewise, the term reactive oxygen species — adopted<br />

at some stage in order to highlight non-radical oxidants such as H 2O 2 and 1 O 2<br />

— now fails to reflect the rich variety of other reactive species in free radical<br />

biology and medicine encompassing nitrogen- , sulfur- , oxygen- , and carboncentered<br />

radicals. With the discovery of nitric oxide, nitrogen-centered radicals<br />

xi


xii Oxidative Stress and Inflammatory Mechanisms<br />

gathered momentum and have matured into an area of enormous importance in<br />

biology and medicine. Nitric oxide or nitrogen monoxide (NO), a free radical<br />

generated in a variety of cell types by nitric oxide synthases (NOSs), is involved<br />

in a wide array of physiological and pathophysiological phenomena such as<br />

vasodilation, neuronal signaling, and inflammation. Of great importance is the<br />

radical–radical reaction of nitric oxide with superoxide anion. This is among<br />

the most rapid non-enzymatic reactions in biology (well over the diffusioncontrolled<br />

limits) and yields the potent non-radical oxidant, peroxynitrite. The<br />

involvement of this species in tissue injury through oxidation and nitration<br />

reactions is well documented.<br />

Virtually all diseases thus far examined involve free radicals. In most cases,<br />

free radicals are secondary to the disease process, but in some instances causality<br />

is established by free radicals. Thus, there is a delicate balance between oxidants<br />

and antioxidants in health and disease. Their proper balance is essential for<br />

ensuring healthy aging.<br />

Both reactive oxygen and nitrogen species are involved in the redox regulation<br />

of cell functions. Oxidative stress is increasingly viewed as a major upstream<br />

component in the signaling cascade involved in inflammatory responses, stimulation<br />

of cell adhesion molecules, and chemoattractant production and as an early<br />

component in age-related neurodegenerative disorders such as Alzheimer’s, Parkinson’s,<br />

and Huntington’s diseases, and amyotrophic lateral sclerosis. Hydrogen<br />

peroxide is probably the most important redox signaling molecule that, among<br />

others, can activate NFκB, Nrf2, and other universal transcription factors. Increasing<br />

steady-state levels of hydrogen peroxide have been linked to a cell’s redox<br />

status with clear involvement in adaptation, proliferation, differentiation, apoptosis,<br />

and necrosis.<br />

The identification of oxidants in regulation of redox cell signaling and gene<br />

expression was a significant breakthrough in the field of oxidative stress: the<br />

classical definition of oxidative stress as an imbalance between the production<br />

of oxidants and the occurrence of cell antioxidant defenses proposed by Sies<br />

in 1985 now seems to provide a limited concept of oxidative stress, but it<br />

emphasizes the significance of cell redox status. Because individual signaling<br />

and control events occur through discrete redox pathways rather than through<br />

global balances, a new definition of oxidative stress was advanced by Dean P.<br />

Jones (Antioxidants & Redox Signaling [2006]) as a disruption of redox signaling<br />

and control that recognizes the occurrence of compartmentalized cellular<br />

redox circuits. Recognition of discrete thiol redox circuits led Jones to provide<br />

this new definition of oxidative stress. Measurements of GSH/GSSG, cysteine/cystine,<br />

or thioredoxin reduced/thioredoxin oxidized provide a quantitative definition<br />

of oxidative stress. Redox status is thus dependent on the degree to which<br />

tissue-specific cell components are in the oxidized state.<br />

In general, the reducing environments inside cells help to prevent oxidative<br />

damage. In this reducing environment, disulfide bonds (S–S) do not spontaneously<br />

form because sulfhydryl groups are maintained in the reduced state (SH), thus<br />

preventing protein misfolding or aggregation. The reducing environment is


Series Preface xiii<br />

maintained by metabolism and by the enzymes involved in maintenance of<br />

thiol/disulfide balance and substances such as glutathione, thioredoxin, vitamins<br />

E and C, and enzymes such as superoxide dismutases, catalase, and the seleniumdependent<br />

glutathione reductase and glutathione and thioredoxin-dependent<br />

hydroperoxidases (periredoxins) that serve to remove reactive oxygen species<br />

(hydroperoxides). Also of importance is the existence of many tissue- and cell<br />

compartment-specific isoforms of antioxidant enzymes and proteins.<br />

Compelling support for the involvement of free radicals in disease development<br />

originates from epidemiological studies showing that enhanced antioxidant<br />

status is associated with reduced risk of several diseases. Of great significance<br />

is the role that micronutrients play in modulation of redox cell signaling; this<br />

establishes a strong linking of diet and health and disease centered on the abilities<br />

of micronutrients to regulate redox cell signaling and modify gene expression.<br />

These concepts are anticipated to serve as platforms for the development of<br />

tissue-specific therapeutics tailored to discrete, compartmentalized redox circuits.<br />

This, in essence, dictates principles of drug development-guided knowledge of<br />

mechanisms of oxidative stress. Hence, successful interventions will take advantage<br />

of new knowledge of compartmentalized redox control and free radical<br />

scavenging.<br />

<strong>OXIDATIVE</strong> <strong>STRESS</strong> IN HEALTH <strong>AND</strong> <strong>DISEASE</strong><br />

Oxidative stress is an underlying factor in health and disease. In this series of<br />

books, the importance of oxidative stress and diseases associated with organ<br />

systems of the body is highlighted by exploring the scientific evidence and clinical<br />

applications of this knowledge. This series is intended for researchers in the basic<br />

biomedical sciences and clinicians. The potential of such knowledge for healthy<br />

aging and disease prevention warrants further knowledge about how oxidants and<br />

antioxidants modulate cell and tissue function.<br />

Lester Packer<br />

Enrique Cadenas


Preface<br />

Metabolic syndrome is a multicomponent disorder characterized by obesity, insulin<br />

resistance, dyslipidemia, and hypertension that is associated with the risks of<br />

type 2 diabetes mellitus and cardiovascular disease. Obesity plays a central role<br />

and is a principal causative factor in the progression of metabolic syndrome<br />

leading to type 2 diabetes mellitus and cardiovascular disease. Increased systemic<br />

oxidative stress may be an important mechanism by which obesity increases the<br />

incidence of atherosclerotic cardiovascular disease, the progression of which<br />

entails oxidative (up-regulation of oxidative enzymes) and inflammatory<br />

(increased production of tumor necrosis factors and other cytokines) components.<br />

The significance of inflammatory processes in accumulated fat appears to be an<br />

early initiator of metabolic syndrome, thus strengthening an association with<br />

oxidative stress and supporting the development of drugs targeted at ameliorating<br />

cellular redox status. Likewise, the more active angiotensin system in obesity<br />

may contribute to expanded oxidative stress that serves as a key signaling event<br />

in vascular remodeling.<br />

A greater understanding of the molecular mechanisms that underlie inflammation<br />

and oxidative stress with implications for metabolic stress must be<br />

achieved so that evidence-based nutritional and pharmacological therapies can<br />

be developed to attenuate the impacts of obesity-induced insulin resistance and<br />

ensuing metabolic syndrome.<br />

The chapters in this book report cutting-edge research exploring the intracellular<br />

events mediating or preventing oxidative stress and pro-inflammatory processes<br />

in obesity and type 2 diabetes as well as the molecular mechanisms inherent<br />

in the progression of metabolic stress, phenotypic perspectives, dietary factors,<br />

and micronutrients. They attempt to provide perspectives on the different components<br />

of metabolic stress and obesity and their associations with oxidative stress<br />

and inflammation. Pharmacological interventions in disease management are not<br />

emphasized and are covered elsewhere.<br />

The editors gratefully appreciate the initial stimulus for this book — a workshop<br />

on obesity, oxidative stress related to metabolic syndrome, uncoupling<br />

proteins, and micronutrient action conducted March 15 through 18, 2006 as part<br />

of the Twelfth Annual Meeting of the Oxygen Club of California at Santa Barbara.<br />

We would like to acknowledge the sponsorship of that workshop by the Human<br />

Nutrition Group of BASF, Ludwigshafen, Germany, and, in particular, the input<br />

and help of Dr. Ute Obermüller-Jevic.<br />

xv


About the Editors<br />

Lester Packer received his PhD in microbiology and biochemistry from Yale<br />

University. For many years he was professor and senior researcher at the University<br />

of California at Berkeley. Currently he is an adjunct professor in the<br />

Department of Pharmacology and Pharmaceutical Sciences at the University of<br />

Southern California. Recently, he was appointed distinguished professor at the<br />

Institute of Nutritional Sciences of the Chinese Academy of Sciences, Shanghai,<br />

China. His research interests are related to the molecular, cellular, and physiological<br />

role of oxidants, free radicals, antioxidants, and redox regulation in health<br />

and disease.<br />

Professor Packer is the recipient of numerous scientific achievement awards<br />

including three honorary doctoral degrees. He has served as president of the<br />

International Society of Free Radical Research (SRRRI), president of the Oxygen<br />

Club of California (OCC), and vice-president of UNESCO—the United Nations<br />

Global Network on Molecular and Cell Biology (MCBN).<br />

Helmut Sies earned an MD from the University of Munich, Germany, and an<br />

honorary PhD from the University of Buenos Aires, Argentina. He is a professor<br />

and chairman at the Institute of Biochemistry and Molecular Biology I at Heinrich<br />

Heine University at Dusseldorf, Germany. He served as a visiting professor at<br />

the University of California (Berkeley) and is an adjunct professor at the University<br />

of Southern California. He is a fellow of the National Foundation for<br />

Cancer Research based in Bethesda, Maryland and a fellow of the Royal College<br />

of Physicians of London.<br />

Dr. Sies has been president of the Society for Free Radical Research International<br />

and is also the president of the Oxygen Club of California. His research<br />

interests focus on the field of oxidative stress, oxidants, and antioxidants.<br />

xvii


Contributors<br />

Farhad Amiri, PhD<br />

Lady Davis Institute for Medical<br />

Research<br />

Sir Mortimer B. Davis-Jewish General<br />

Hospital<br />

McGill University<br />

Montreal, Canada<br />

Anna Aronis<br />

School of Nutritional Sciences<br />

Institute of Biochemistry, Food<br />

Science, and Nutrition<br />

The Hebrew University of Jerusalem<br />

Rehovot, Israel<br />

Karim Benkirane, PhD<br />

Lady Davis Institute for Medical<br />

Research<br />

Sir Mortimer B. Davis-Jewish General<br />

Hospital<br />

McGill University<br />

Montreal, Canada<br />

Jennifer Berliner, MD<br />

Cardiology Section<br />

Feinberg School of Medicine<br />

Northwestern University<br />

Chicago, Illinois<br />

T. Bobbert<br />

Abteilung fur Endokrinologie,<br />

Diabetes und Ernahrung<br />

Charite Universitatsmedizin<br />

Berlin, Germany<br />

Joshua D. Brooks<br />

Division of Clinical Pharmacology<br />

Vanderbilt University Medical<br />

Center<br />

Nashville, Tennessee<br />

Antonio Ceriello, MD<br />

Warwick Medical School<br />

University of Warwick<br />

Coventry, United Kingdom<br />

Catherine B. Chan<br />

Department of Biomedical<br />

Sciences<br />

Atlantic Veterinary College<br />

University of Prince Edward Island<br />

Charlottetown, Canada<br />

Ajay Chaudhuri, MD<br />

School of Medicine and Biomedical<br />

Sciences<br />

State University of New York<br />

and<br />

Kaleida Health<br />

Buffalo, New York<br />

An-Sik Chung<br />

Department of Biological<br />

Sciences<br />

Korea Advanced Institute of Science<br />

and Technology<br />

Daejeon, Republic of Korea<br />

xix


xx Oxidative Stress and Inflammatory Mechanisms<br />

Paresh Dandona, MD, PhD<br />

School of Medicine and Biomedical<br />

Sciences<br />

State University of New York<br />

and<br />

Kaleida Health<br />

Buffalo, New York<br />

Sridevi Devaraj<br />

Laboratory for Atherosclerosis and<br />

Metabolic Research<br />

University of California at Davis<br />

Medical Center<br />

Sacramento, California<br />

Jingyu Diao<br />

Department of Physiology<br />

Faculty of Medicine<br />

University of Toronto<br />

Toronto, Canada<br />

John Alan Farmer<br />

Baylor College of Medicine<br />

Houston, Texas<br />

Ling Gao<br />

Division of Clinical Pharmacology<br />

Vanderbilt University Medical Center<br />

Nashville, Tennessee<br />

Husam Ghanim, PhD<br />

School of Medicine and Biomedical<br />

Sciences<br />

State University of New York<br />

and<br />

Kaleida Health<br />

Buffalo, New York<br />

Barry J. Goldstein<br />

Dorrance Hamilton Research<br />

Laboratories<br />

Jefferson Medical College<br />

Thomas Jefferson University<br />

Philadelphia, Pennsylvania<br />

David Heber, MD, PhD, FACP,<br />

FACN<br />

Center for Human Nutrition<br />

David Geffen School of Medicine<br />

University of California<br />

Los Angeles, California<br />

Erik J. Henriksen<br />

Department of Physiology<br />

College of Medicine<br />

University of Arizona<br />

Tucson, Arizona<br />

Sushil K. Jain<br />

Department of Pediatrics<br />

Health Sciences Center<br />

Louisiana State University<br />

Shreveport, Louisiana<br />

Ishwarlal Jialal, MD, PhD<br />

Laboratory for Atherosclerosis and<br />

Metabolic Research<br />

University of California at Davis<br />

Medical Center<br />

Sacramento, California<br />

Janet C. King<br />

Children’s Hospital<br />

Oakland Research Institute<br />

Oakland, California<br />

Berthold Koletzko, MD<br />

Division of Metabolic Disorders and<br />

Nutrition<br />

Dr. von Hauner Children’s Hospital<br />

Ludwig Maximilians University<br />

Munich, Germany<br />

Ki-Up Lee<br />

Department of Internal Medicine<br />

University of Ulsan College of<br />

Medicine<br />

Seoul, Republic of Korea


Contributors xxi<br />

Woo Je Lee<br />

Department of Internal Medicine<br />

Inje University College of Medicine<br />

Seoul, Republic of Korea<br />

Kalyankar Mahadev<br />

Dorrance Hamilton Research<br />

Laboratories<br />

Jefferson Medical College<br />

Thomas Jefferson University<br />

Philadelphia, Pennsylvania<br />

Ginger L. Milne<br />

Division of Clinical Pharmacology<br />

Vanderbilt University Medical<br />

Center<br />

Nashville, Tennessee<br />

Paul Mitrani<br />

Department of Biochemistry<br />

School of Medicine and Biomedical<br />

Sciences<br />

State University of New York<br />

Buffalo, New York<br />

Priya Mohanty, MD<br />

School of Medicine and Biomedical<br />

Sciences<br />

State University of New York<br />

and<br />

Kaleida Health<br />

Buffalo, New York<br />

Jason D. Morrow, MD<br />

Division of Clinical Pharmacology<br />

Vanderbilt University Medical<br />

Center<br />

Nashville, Tennessee<br />

Jong-Min Park<br />

Department of Biological Sciences<br />

Korea Advanced Institute of Science<br />

and Technology<br />

Daejeon, Republic of Korea<br />

Joong-Yeol Park<br />

Department of Internal Medicine<br />

University of Ulsan College of<br />

Medicine<br />

Seoul, Republic of Korea<br />

Mulchand S. Patel<br />

Department of Biochemistry<br />

School of Medicine and Biomedical<br />

Sciences<br />

State University of New York<br />

Buffalo, New York<br />

Helen M. Roche<br />

Institute of Molecular Medicine<br />

Trinity Health Sciences Centre<br />

St. James Hospital<br />

Dublin, Republic of Ireland<br />

Ernesto L. Schiffrin, MD, PhD,<br />

FRSC, FRCPC, FACP<br />

Lady Davis Institute for Medical<br />

Research<br />

Sir Mortimer B. Davis-Jewish General<br />

Hospital<br />

McGill University<br />

Montreal, Canada<br />

Hiromichi Shoji<br />

Department of Pediatrics<br />

Juntendo University School of<br />

Medicine<br />

Tokyo, Japan<br />

and<br />

Division of Metabolic Disorders and<br />

Nutrition<br />

Dr. von Hauner Children’s<br />

Hospital<br />

Ludwig Maximilians University<br />

Munich, Germany


xxii Oxidative Stress and Inflammatory Mechanisms<br />

Uma Singh<br />

Laboratory for Atherosclerosis and<br />

Metabolic Research<br />

University of California at Davis<br />

Medical Center<br />

Sacramento, California<br />

Joachim Spranger<br />

Department of Clinical Nutrition<br />

German Institute of Human Nutrition<br />

Potsdam-Rehbruecke<br />

Nuthetal, Germany<br />

Malathi Srinivasan<br />

Department of Biochemistry<br />

School of Medicine and Biomedical<br />

Sciences<br />

State University of New York<br />

Buffalo, New York<br />

Neil J. Stone, MD<br />

Cardiology Section<br />

Feinberg School of Medicine<br />

Northwestern University<br />

Chicago, Illinois<br />

Oren Tirosh<br />

School of Nutritional Sciences<br />

Institute of Biochemistry, Food<br />

Science, and Nutrition<br />

The Hebrew University of Jerusalem<br />

Rehovot, Israel<br />

Michael B. Wheeler<br />

Department of Physiology<br />

Faculty of Medicine<br />

University of Toronto<br />

Toronto, Canada<br />

Xiangdong Wu<br />

Dorrance Hamilton Research<br />

Laboratories<br />

Jefferson Medical College<br />

Thomas Jefferson University<br />

Philadelphia, Pennsylvania<br />

Yuichiro Yamashiro<br />

Department of Pediatrics<br />

Juntendo University School of<br />

Medicine<br />

Tokyo, Japan


Section I<br />

Oxidative Stress, Metabolic<br />

Syndrome, Obesity, Diabetes,<br />

and Uncoupling Proteins


1<br />

CONTENTS<br />

The Metabolic<br />

Syndrome Defined<br />

Neil J. Stone and Jennifer Berliner<br />

Rationale for Defining Metabolic Syndrome .......................................................3<br />

Metabolic Syndrome Definitions..........................................................................5<br />

World Health Organization (WHO) Definition...........................................5<br />

ATP III Initial Report ..................................................................................5<br />

American Association of Clinical Endocrinologists Guidelines ................5<br />

European Group for Study of Insulin Resistance Syndrome .....................7<br />

Revised ATP III Guidelines.........................................................................7<br />

International Diabetes Federation (IDF) Definition ...................................8<br />

Is Metabolic Syndrome Useful and Which Definition Should be Used?............9<br />

References ...........................................................................................................13<br />

RATIONALE FOR DEFINING METABOLIC SYNDROME<br />

In the past decade, sedentary lifestyles, atherogenic high calorie diets, and weight<br />

gains have characterized adolescents and adults in the United States and in many<br />

countries across the globe. 1,2 Indeed, a recent report 2 estimated the prevalences<br />

of overweight and obese people in the U.S. above 60 and 30%, respectively. This<br />

is not a unique burden for the U.S., but reflects a worldwide trend demonstrating<br />

an increased prevalence in metabolic risk factors 3,4 including visceral obesity,<br />

insulin resistance, dyslipidemia with abnormal values for triglycerides and high<br />

density lipoprotein cholesterol (HDL-c), hypertension, and (if measured) prothrombotic<br />

and inflammatory markers.<br />

Although metabolic risk factors track with body mass index (BMI) calculated<br />

as weight/height squared, this has not proven the best correlate of coronary heart<br />

disease (CHD) events. More than 50 years ago, a French investigator drew<br />

attention to fat distribution in a graphic way showing the difference between those<br />

with android or abdominal patterns and those with gynecoid or female patterns<br />

of fat distribution. 5 In 2004, the global INTERHEART project demonstrated that<br />

the measure of obesity most directly related to myocardial infarction (MI) was<br />

3


4 Oxidative Stress and Inflammatory Mechanisms<br />

not BMI, but a measure of body fatness, the waist:hip ratio. 6 Although not as<br />

good as the waist:hip ratio, waist measurements were more predictive than BMI.<br />

In addition to the appreciation that weight gain and obesity are increasing<br />

global problems and that the location of body fat has both prognostic and therapeutic<br />

importance is the recognition of atherogenic risk factor clustering. This<br />

concept was noted by Framingham investigators who found it was common in<br />

both men and women, worsened with weight gain, and appreciably increased the<br />

risk of CHD. 7 The odds ratios for individuals with this risk factor clustering was<br />

more than two-fold for men and more than six-fold for women. Subsequent<br />

studies identified factors that can intensify these metabolic abnormalities. They<br />

include aging (higher prevalence at higher age), ethnic subgroups, endocrine<br />

dysfunction (increased prevalence in polycystic ovary syndrome), and physical<br />

inactivity. 8 The distinction from those unaffected is not trivial. Characteristic<br />

clinical features of risk factor clustering identify individuals who are more likely<br />

to become diabetic and/or experience a cardiac event. This spurred interest in<br />

viewing this constellation of clustered metabolic risk factors as a syndrome. It is<br />

important to recognize that metabolic syndrome is not a disease entity with a<br />

single etiologic cause. Nonetheless, the metabolic syndrome designation seems<br />

apt if you define a syndrome as “a set of symptoms or conditions that occur<br />

together and suggest specific underlying factors, prognosis, or guide to treatment.”<br />

Like the clinical syndrome known as heart failure, there may be great utility<br />

in recognizing patients who meet diagnostic criteria. Although a consensus has<br />

not emerged, current definitions identify patients with metabolic risk factors that<br />

can be seen over their lifespans, from an early stage when risk factors are<br />

emerging until they are fully established later. Although definitions vary, some<br />

believe metabolic syndrome is a useful construct even for those who have developed<br />

diabetes. Clinicians are increasingly recognizing metabolic syndrome<br />

among those with CHD and stroke. Accordingly, this syndrome may be progressive<br />

in nature. Insulin resistance, along with weight gain and obesity, is a major<br />

underlying risk factor for this syndrome, but as will be seen, most definitions do<br />

not consider metabolic syndrome strictly synonymous with insulin resistance.<br />

Moreover, although metabolic syndrome predicts CHD, it is not a replacement<br />

for Framingham risk scoring which, according to the ATP III panel, is required<br />

for setting LDL-c goals to determine intensity of treatment. 4 In clinical practice,<br />

recognition of metabolic syndrome is a useful tool for helping both patients and<br />

clinical care teams understand the adverse prognosis for both diabetes and CHD<br />

and direct therapeutic interventions. Furthermore, those with metabolic syndrome<br />

face increased risks for CHD that exceed the sum of the risk factors. 8 This should<br />

not be surprising given the systemic factors such as inflammation and hypercoagulability<br />

that accompany metabolic syndrome and its attendant visceral obesity.<br />

3 Most important, several well controlled clinical trials have established that<br />

in individuals at high risk for progression to diabetes (and most likely with<br />

metabolic syndrome), a therapeutic lifestyle regimen directed at diet, regular<br />

exercise, and modest weight loss reduces the progression to type 2 diabetes by<br />

almost 60%. 9,10


The Metabolic Syndrome Defined 5<br />

Next we will consider various published definitions of metabolic syndrome<br />

in the hope that it will be possible to gain perspective on the merits and potential<br />

shortcomings of each definition.<br />

METABOLIC SYNDROME DEFINITIONS<br />

WORLD HEALTH ORGANIZATION (WHO) DEFINITION<br />

In 1998, the WHO proposed a set of criteria 11 to define metabolic syndrome. Its<br />

definition required the presence of insulin resistance as a component of the<br />

diagnosis. Insulin resistance was defined as the diagnosis of type 2 diabetes<br />

mellitus, impaired fasting glucose, impaired glucose tolerance or, for those with<br />

normal fasting glucose levels (30 kg/m 2 , and urinary albumin excretion rate<br />

>20 μg/min (Table 1.1).<br />

ATP III INITIAL REPORT<br />

In 2001, the National Cholesterol Education Program (NCEP) introduced definitions<br />

of metabolic syndrome, a constellation of major risk factors, life-habit risk<br />

factors, and emerging risk factors for CHD in its guidelines. The ATP III Expert<br />

Panel 12 recognized metabolic syndrome as a secondary target of risk reduction<br />

therapy, after the primary target of LDL cholesterol was attained. Its definition<br />

of metabolic syndrome included abdominal obesity, triglyceride levels, HDL<br />

cholesterol, blood pressure, and fasting glucose. The diagnosis of metabolic<br />

syndrome was made when three or more of the risk determinants were present<br />

(Table 1.2).<br />

AMERICAN ASSOCIATION OF CLINICAL ENDOCRINOLOGISTS GUIDELINES<br />

The guidelines of the American Association of Clinical Endocrinologists 13 appear<br />

to be a hybrid between the ATP III and WHO definitions of metabolic syndrome,<br />

but no defined number of risk factors is specified. Diagnosis is left to clinical<br />

judgment. The criteria include obesity, elevated triglycerides, low HDL cholesterol,<br />

elevated blood pressure, elevated fasting glucose and other risk factors,<br />

including polycystic ovary syndrome, sedentary lifestyle, advancing age, belonging<br />

to an ethnic group at high risk for type 2 diabetes or cardiovascular disease,<br />

family history of type 2 diabetes, hypertension, and cardiovascular disease<br />

(Table 1.3).


6 Oxidative Stress and Inflammatory Mechanisms<br />

TABLE 1.1<br />

WHO Diagnostic Criteria for Metabolic Syndrome<br />

Component Criteria Comments<br />

Insulin resistance as identified<br />

by type 2 diabetes, impaired<br />

fasting glucose, or impaired<br />

glucose tolerance<br />

Any two of the following:<br />

body mass index; waist:hip ratio >30 kg/m 2 ; >0.9 in men or<br />

>0.85 in women<br />

For those with normal fasting<br />

glucose levels (88 cm (>35 inches)<br />

Triglycerides ≥150 mg/dL<br />

High-density lipoprotein cholesterol Men:


The Metabolic Syndrome Defined 7<br />

TABLE 1.3<br />

American Association of Clinical Endocrinologists Diagnostic Criteria<br />

for Insulin Resistance Syndrome<br />

Risk Factor Defining Level<br />

Elevated triglycerides >150 mg/dL<br />

Low HDL Men: 140 mg/dL<br />

Fasting glucose 110 to 125 mg/dL<br />

Other risk factors Family history of type 2 diabetes, hypertension, or<br />

cardiovascular disease; history of CVD, HTN, NAFLD,<br />

aeonthosis nigricans, history of glucose intolerance or<br />

gestational dialoetes mellitus, BMI >25 kg/m2, polycystic<br />

ovary syndrome; sedentary lifestyle; >40 age, membership<br />

in ethnic group having high risk for type 2 diabetes mellitus<br />

or cardiovascular disease<br />

Source: From Einhorn, D. et al., Endocr Pract 9, 237, 2003.<br />

EUROPEAN GROUP FOR STUDY OF INSULIN RESISTANCE SYNDROME<br />

In 1999, the European Group for Study of Insulin Resistance (EGIR) proposed<br />

a modification of the WHO definition. 14 This group used the term insulin resistance<br />

syndrome rather than metabolic syndrome. They likewise assumed that<br />

insulin resistance is the major cause, and required evidence of it for diagnosis.<br />

EGIR defines insulin resistance syndrome as the presence of insulin resistance<br />

or fasting hyperinsulinemia (highest 25%) and two of the following: hyperglycemia,<br />

hypertension, dyslipidemia, and central obesity. All these criteria must be<br />

measured before it is possible to evaluate the presence of the syndrome. Hyperglycemia<br />

should be defined at fasting values to provide reproducible criteria<br />

(Table 1.4).<br />

REVISED ATP III GUIDELINES<br />

In an updated version of the original ATP III guidelines, 15 the threshold for<br />

impaired fasting glucose (IFG) was reduced from ≥110 to ≥100 mg/dL. This<br />

adjustment corresponded to the recently modified American Diabetes Association<br />

(ADA) criteria for IFG. The rest of the criteria were essentially the same although<br />

several points were clarified. A diagnosis of the metabolic syndrome required<br />

three of the following diagnoses: abdominal obesity (as waist circumference),<br />

elevated triglycerides, reduced HDL cholesterol, hypertension, and elevated fasting<br />

glucose. As noted in ATP III, some people will manifest features of insulin<br />

resistance and metabolic syndrome with only moderate increases in waist


8 Oxidative Stress and Inflammatory Mechanisms<br />

TABLE 1.4<br />

European Group for Study of Insulin Resistance Syndrome Criteria<br />

Risk Factor Defining Level<br />

Defined by presence of insulin resistance<br />

or fasting hyperinsulinemia (highest<br />

25%) <strong>AND</strong> two of the following:*<br />

Hyperglycemia Fasting plasma glucose ≥6.1 mmol/L, but nondiabetic<br />

Hypertension Systolic/diastolic blood pressure ≥140/90 mm Hg or<br />

treated for hypertension<br />

Dyslipidemia Triglycerides >2.0 mmol/L or HDL-cholesterol 3 mg/L, microalbuminuria,<br />

impaired glucose tolerance, and elevated total apolipoprotein B.<br />

Some populations are predisposed to insulin resistance, metabolic syndrome,<br />

and type 2 diabetes mellitus, with only moderate increases in waist circumference<br />

(i.e., populations from South Asia, China, Japan, and other Asian countries). None<br />

of these phenotypic features or ethnic differences was included in the ATP III<br />

diagnostic criteria, but if individuals with such characteristics have only moderate<br />

elevations of waist circumference plus at least two ATP III metabolic syndrome<br />

features, the writing group suggests that consideration should be given to managing<br />

them in the same way people with three ATP III risk factors are managed<br />

(Table 1.5).<br />

INTERNATIONAL DIABETES FEDERATION (IDF) DEFINITION<br />

The IDF clinical definition 16 requires the presence of abdominal obesity for<br />

diagnosis. When abdominal obesity is present, two additional factors originally<br />

listed in the ATP III definition are sufficient for diagnosis: raised concentration<br />

of triglycerides, reduced concentration of HDL cholesterol, hypertension, and<br />

raised fasting plasma glucose concentration or previously diagnosed type 2 diabetes.<br />

IDF recognized and emphasized ethnic differences in the correlation of<br />

abdominal obesity and other metabolic syndrome risk factors. For this reason,<br />

criteria of abdominal obesity were specified by nationality or ethnicity based on


The Metabolic Syndrome Defined 9<br />

TABLE 1.5<br />

Revised ATP III Criteria<br />

Risk Factor<br />

(Any Three for Diagnosis) Defining Level<br />

Elevated waist circumference* Men: ≥102 cm (≥40 inches)<br />

Women: ≥88 cm (≥35 inches)<br />

Elevated triglycerides >150 mg/dL (1.7 mmol/L) or on drug treatment for elevated TG<br />

Reduced HDL cholesterol Men:


10 Oxidative Stress and Inflammatory Mechanisms<br />

TABLE 1.6<br />

International Diabetes Federation Criteria for Metabolic Syndrome<br />

Risk Factor Defining Level<br />

Abdominal obesity* <strong>AND</strong> two<br />

additional factors:<br />

Triglycerides ≥150 mg/dL or specific treatment for this abnormality<br />

Reduced concentration of HDL


The Metabolic Syndrome Defined 11<br />

lipids, diabetes, smoking, hypertension, metabolic and lifestyle factors such as<br />

abdominal obesity, psychosocial factors, levels of consumption of fruits and<br />

vegetables, alcohol intake, and regular physical activity.<br />

Interestingly, these investigators created a study definition of metabolic<br />

syndrome by using self-reported diabetes and hypertension as surrogates for<br />

measured fasting blood sugar and blood pressure, measured apo B and apo A1<br />

as surrogates for fasting triglycerides and HDL-c, and measured waist:hip ratio<br />

instead of waist circumference. Not surprisingly they found that their dichotomous<br />

metabolic syndrome variables were not as good as predictors of acute MI<br />

as the nine variables listed above that predicted over 90% of cases of acute MI.<br />

This has been a criticism of the metabolic syndrome, but in fairness, ATP III<br />

always emphasized that patients (especially those having two or more risk factors)<br />

need Framingham risk scoring with its continuous variables to calculate global<br />

risk and suggested attention to metabolic variables after LDL-c goals had been<br />

realized. For example, the ATP III algorithm notes that if triglycerides are 200<br />

mg/dL or more despite attainment of the LDL-c goal, then treatment to achieve<br />

non-HDL-c goals (some may prefer apo B) is recommended.<br />

It is hoped that with further investigation, consensus will eventually be<br />

reached on a single definition for metabolic syndrome.<br />

Table 1.8 summarizes unique features, drawbacks, and comments regarding<br />

the definitions discussed in this chapter. Endocrinologists and investigators may<br />

be drawn to definitions that focus on insulin resistance. Clinicians may prefer the<br />

easily identified clinical features of ATP III that improve with lifestyle changes<br />

and the resultant modest degrees of weight loss. Further studies and discussion<br />

hopefully will clarify the final form of the metabolic syndrome definition. Until<br />

then, regardless of the criteria utilized, a systematic identification of metabolic<br />

variables leads invariably to an appropriate strong focus on the need for improved<br />

diet, regular physical activity, and weight reduction that most would agree is an<br />

important first step in the comprehensive approach to reducing the burden of<br />

diabetes and CHD worldwide.


12 Oxidative Stress and Inflammatory Mechanisms<br />

TABLE 1.8<br />

Features, Drawbacks, and Comments on Various Definitions of Metabolic<br />

Syndrome<br />

Defining<br />

Organization Unique Features Drawbacks Comments<br />

WHO Requires insulin resistance;<br />

BP requirements higher than<br />

ATP III; HDL cholesterol<br />

requirements lower; uses<br />

BMI instead of waist<br />

circumference;<br />

microalbuminuria is<br />

criterion<br />

2001 ATP III<br />

and update<br />

Uses abdominal obesity<br />

rather than BMI as criterion;<br />

simple criteria for diagnosis;<br />

update clarified several<br />

diagnosis issues<br />

AACE Uses BMI rather than<br />

abdominal obesity as<br />

criterion<br />

EGIR Requires insulin resistance<br />

for diagnosis; excludes<br />

patients with type 2 diabetes<br />

mellitus; BP requirements<br />

higher than ATP III; uses<br />

abdominal obesity rather<br />

than BMI as criterion;<br />

simple criteria for diagnosis<br />

IDF Requires abdominal obesity<br />

for diagnosis; simple criteria<br />

(same as ATP III) for<br />

diagnosis<br />

Requires glucose<br />

testing that may<br />

require a separate<br />

office or hospital<br />

visit<br />

No measure of<br />

insulin resistance;<br />

this would detract<br />

from ease of use of<br />

these criteria<br />

Reproducibility; no<br />

defined number of<br />

risk factors<br />

specified<br />

Insulin levels not<br />

reliable in clinical<br />

practice setting<br />

Physician must have<br />

data on waist<br />

circumference for<br />

ethnic subgroups<br />

Demonstrating insulin<br />

resistance in a patient without<br />

type 2 diabetes usually<br />

involves oral glucose<br />

tolerance testing or<br />

hyperinsulinemic/euglycemic<br />

clamp testing; considered<br />

costly or inconvenient in<br />

many clinical practices<br />

Weight loss achieved by better<br />

diet and regular exercise has<br />

potential to improve all<br />

components of the metabolic<br />

syndrome definition<br />

Diagnosis left to clinical<br />

judgment; this makes<br />

estimates of prognosis from<br />

clinical studies difficult<br />

Unlike some definitions, it<br />

excludes type 2 diabetes<br />

mellitus; some would argue<br />

that therapeutic implications<br />

of metabolic syndrome may<br />

be valuable in diabetics with<br />

increased waist<br />

circumferences<br />

Uses a recognizable and easily<br />

measurable clinical feature<br />

(increased waist<br />

circumference) to initiate<br />

consideration of diagnosis


The Metabolic Syndrome Defined 13<br />

REFERENCES<br />

1. Janssen, I. et al. Health Behaviour in School-Aged Children Obesity Working<br />

Group: comparison of overweight and obesity prevalence in school-aged youth<br />

from 34 countries and their relationships with physical activity and dietary patterns.<br />

Obes Rev 6, 123, 2005.<br />

2. Flegal, K.M. et al. Prevalence and trends in obesity among U.S. adults, 1999–2000.<br />

JAMA 288, 1723, 2002.<br />

3. Eckel, R., Grundy, S., and Zimmet, P. The metabolic syndrome. Lancet 365, 1415,<br />

2005.<br />

4. Grundy, S.M. Metabolic syndrome scientific statement by the American Heart<br />

Association and the National Heart, Lung, and Blood Institute. Arterioscler<br />

Thromb Vasc Biol 25, 2243, 2005.<br />

5. Vague, J. La differenciation sexuelle, facteur determinant des formes de l’obesite.<br />

Presse Med 30, 339, 1947.<br />

6. Yusuf, S. et al. INTERHEART Study Investigators: obesity and the risk of myocardial<br />

infarction in 27,000 participants from 52 countries: a case-control study.<br />

Lancet 366, 1640, 2005.<br />

7. Wilson, P.F.F., Kannel, W.B., Silbershatz, H., and D’Agostino, R.B. Clustering of<br />

metabolic risk factors and coronary heart disease. Arch Int Med 159, 1104, 1999.<br />

8. Grundy, S. Metabolic syndrome: connecting and reconciling cardiovascular and<br />

diabetes worlds. JACC 47, 1093, 2006.<br />

9. Diabetes Prevention Program Research Group. Reduction in the incidence of type<br />

2 diabetes with lifestyle intervention or metformin. New Engl J Med 346, 393,<br />

2002.<br />

10. Tuomilehto, J. et al. Finnish Diabetes Prevention Study Group: prevention of type<br />

2 diabetes mellitus by changes in lifestyle among subjects with impaired glucose<br />

tolerance. New Engl J Med 344, 1343, 2001.<br />

11. Alberti, K.G. and Zimmet, P.Z. Definition, diagnosis and classification of diabetes<br />

mellitus and its complications. Part 1. Diagnosis and classification of diabetes<br />

mellitus provisional report of a WHO consultation. Diabet Med 15, 539, 1998.<br />

12. Adult Treatment Panel (ATP) III. Executive summary of third report of the<br />

National Cholesterol Education Program (NCEP) Expert Panel on Detection,<br />

Evaluation, and Treatment of High Blood Cholesterol in Adults. JAMA 285, 2486,<br />

2001.<br />

13. Einhorn, D. et al. American College of Endocrinology position statement on the<br />

insulin resistance syndrome. Endocr Pract 9, 237, 2003.<br />

14. Balkau, B. and Charles, M.A. Comment on the provisional report from the WHO<br />

consultations, European Group for the Study of Insulin Resistance (EGIR). Diabet<br />

Med 16, 442, 1999.<br />

15. Grundy, S.M. et al. Diagnosis and management of the metabolic syndrome: an<br />

American Heart Association/National Heart, Lung, and Blood Institute scientific<br />

statement. Circulation 112, 2735, 2005.<br />

16. Alberti, K.B., Zimmet, P., and Shaw, J. The metabolic syndrome: a new worldwide<br />

definition. Lancet 366, 1059, 2005.<br />

17. Kahn, R., Buse, J., Ferrannini, E., and Stern, M. The metabolic syndrome: time<br />

for a critical appraisal. A joint statement from the American Diabetes Association<br />

and the European Association for the Study of Diabetes. Diabet Care 28, 2289,<br />

2005.


14 Oxidative Stress and Inflammatory Mechanisms<br />

18. Yusuf, S. et al. Effect of potentially modifiable risk factors associated with myocardial<br />

infarction in 52 countries (the INTERHEART study): case-control study.<br />

Lancet 364, 937, 2004.


2<br />

CONTENTS<br />

The Metabolic<br />

Syndrome: The Question<br />

of Balance between the<br />

Pro-Inflammatory Effect<br />

of Macronutrients and<br />

the Anti-Inflammatory<br />

Effect of Insulin<br />

Paresh Dandona, Ajay Chaudhuri,<br />

Priya Mohanty, and Husam Ghanim<br />

Introduction .........................................................................................................15<br />

Insulin Resistance Syndrome: Metabolic Perspective........................................16<br />

Novel Non-Metabolic Actions of Insulin ...........................................................17<br />

Obesity and Inflammation...................................................................................20<br />

Insulin Resistance: Inflammatory Hypothesis ....................................................22<br />

Macronutrients and Origin of Inflammation ......................................................22<br />

Conclusion: Inflammation Hypothesis of Metabolic Syndrome........................25<br />

References ...........................................................................................................26<br />

INTRODUCTION<br />

The common occurrence of the combination of obesity, insulin resistance, hypertension,<br />

hypertriglyceridemia, low HDL cholesterol, and hyperinsulinemia was<br />

first described by Reaven as insulin resistance syndrome or metabolic syndrome.<br />

The syndrome was recognized as a pro-atherogenic risk causing coronary heart<br />

disease (CHD). More recently, other features like elevated plasma PAI-1 and CRP<br />

15


16 Oxidative Stress and Inflammatory Mechanisms<br />

have been added to the combination. In view of recent data demonstrating that<br />

insulin exerts an anti-inflammatory effect while macronutrients exert pro-inflammatory<br />

effects, we are in a better position to explain why an insulin-resistant<br />

state such as metabolic syndrome is pro-inflammatory and also explain how it<br />

develops. This focused review discusses the relevance of these recent observations,<br />

puts into perspective the pathogenesis of various features of metabolic<br />

syndrome, and also predicts some features that may be incorporated into it in the<br />

future.<br />

INSULIN RESISTANCE SYNDROME: METABOLIC PERSPECTIVE<br />

Reaven’s original description of the metabolic syndrome consisted of obesity,<br />

insulin resistance, hypertension, impaired glucose tolerance or diabetes, hyperinsulinemia,<br />

and dyslipidemia characterized by elevated triglyceride and low<br />

HDL concentrations. 1 All these features serve as risk factors for atherosclerosis<br />

and thus metabolic syndrome constitutes a significant risk for coronary heart<br />

disease 2–5 (Table 2.1). The features of obesity or overweight and insulin resistance<br />

also provide significant risks for developing type 2 diabetes. 5,6 The risks for CHD<br />

and diabetes with metabolic syndrome are greater than those for simple obesity<br />

alone without insulin resistance and therefore the understanding of the pathogenesis<br />

and, through it, a rational approach to its therapy are of prime importance.<br />

TABLE 2.1<br />

Classic Biological Effects of Insulin and Classic Metabolic Syndrome<br />

Based on Resistance to Metabolic Effects of Insulin<br />

Nutrients Normal Insulin Action Insulin-Resistant State<br />

Carbohydrates Hepatic glucose production<br />

Glucose utilization<br />

Glycogenesis<br />

Lipids Lipolysis<br />

FFA and glycerol<br />

Lipogenesis<br />

HDL<br />

Triglycerides<br />

Proteins Gluconeogenesis<br />

Amino acids<br />

Protein synthesis<br />

Purines Uric acid clearance<br />

Uric acid formation<br />

Hyperglycemia<br />

Hyperinsulinemia<br />

Source: From Dandona, P. et al., Circulation 111, 1448, 2005.<br />

Lipolysis<br />

FFA and glycerol<br />

Hepatic triglyceride and apo B synthesis<br />

Hypertriglyceridemia<br />

HDL<br />

Small dense LDL<br />

Glyconeogenesis<br />

Protein catabolism<br />

Protein synthesis<br />

Hyperuricemia


The Metabolic Syndrome 17<br />

The original conceptualization of this syndrome was based on resistance to<br />

the metabolic actions of insulin. Thus, hyperinsulinemia, glucose intolerance,<br />

type 2 diabetes, hypertriglyceridemia, and low HDL concentrations could be<br />

accounted for by resistance to the actions of insulin on carbohydrate and lipid<br />

metabolism. While the features described above explain the atherogenesis to some<br />

extent, Reaven maintained that hyperinsulinemia contributed to atherogenicity<br />

and thus insulin is atherogenic, leading to CHD and cerebrovascular disease<br />

associated with this syndrome. However, as our understanding of the action of<br />

insulin evolves to comprehensively include recent discoveries, 7 we can better see<br />

that insulin resistance is the basis of most if not all the features of this syndrome.<br />

Obesity probably leads to hypertension through (1) increased vascular tone<br />

created by a reduced bioavailability of nitric oxide (NO) due to increased oxidative<br />

stress, 8 (2) increased asymmetric dimethyl arginine (ADMA) concentrations, 9<br />

(3) increased sympathetic tone, 10 and (4) increased expression of angiotensinogen<br />

by adipose tissue leading to an activation of the renin–angiotensin system. 11 The<br />

last factor requires further critical investigation.<br />

Metabolic syndrome is characterized by low HDL in association with an<br />

elevated triglyceride concentration. This is believed to be due to an increased<br />

triglyceride load in HDL particles that are acted upon by hepatic lipase that<br />

hydrolyzes the triglyceride. The loss of the triglyceride results in a small HDL<br />

particle filtered by the kidney, resulting in decreases in apolipoprotein A (apo A)<br />

and HDL concentrations. Apart from an increase in the loss of apo A, data<br />

demonstrate that insulin may promote apo A gene transcription. 12 Therefore,<br />

insulin-resistant states may be associated with diminished apo A biosynthesis. 13<br />

An increase in plasma free fatty acid (FFA) concentrations plays a key role<br />

in the pathogenesis of insulin resistance through specific actions that block insulin<br />

signal transduction. Increases of plasma FFA concentrations in normal subjects<br />

to levels comparable to those in the obese also resulted in the induction of<br />

oxidative stress, inflammation, and subnormal vascular reactivity along with<br />

insulin resistance. 14 Because resistance to insulin also results in the relative nonsuppression<br />

of adipocyte hormone-sensitive lipase, there is further enhancement<br />

of lipolysis and an increase in FFA concentration, leading to a vicious cycle of<br />

lipolysis, increased FFA, insulin resistance, and inflammation.<br />

Several new features have been added to the syndrome over time. These<br />

include elevated plasminogen activator inhibitor-1 (PAI-1) concentrations and<br />

elevated C-reactive protein (CRP) concentrations. These features were added on<br />

the basis that they were found frequently in association with metabolic syndrome<br />

and no rational explanation indicated why they actually occurred. These features<br />

are probably related to both insulin resistance and obesity. The relationship of<br />

inflammation to obesity and insulin resistance still needs to be explained. 15<br />

NOVEL NON-METABOLIC ACTIONS OF INSULIN<br />

The nonmetabolic actions of insulin are readily explained by the recent observations<br />

that insulin is an anti-inflammatory hormone and that macronutrient intake is


18 Oxidative Stress and Inflammatory Mechanisms<br />

pro-inflammatory. Insulin has been shown to suppress several pro-inflammatory<br />

transcription factors such as nuclear factor kappa-B (NFκB), early growth response-<br />

1 (Egr-1), activator protein-1 (AP-1), and the corresponding genes they regulate<br />

that mediate inflammation. 16,17 An impairment of the action of insulin due to insulin<br />

resistance would thus result in the activation of these pro-inflammatory transcription<br />

factors and an increase in the expression of the corresponding genes.<br />

Insulin has been shown to suppress NFκB binding activity, reactive oxygen<br />

species (ROS) generation, p47 phox expression, increase inhibitor kappa-B (IκB)<br />

expression in mononuclear cells (MNCs), and suppress plasma concentrations of<br />

intracellular adhesion molecule-1 (ICAM-1) and monocyte chemoattractant protein-1<br />

(MCP-1). 16 In addition, insulin suppresses AP-1 and Egr-1, two pro-inflammatory<br />

transcription factors and their respective genes, matrix metalloproteinase-<br />

9 (MMP-9), tissue factor (TF), and PAI-1. 17–19 Thus, insulin exerts comprehensive<br />

anti-inflammatory effects and also has anti-oxidant effects as reflected in the<br />

suppression of ROS generation and p47 phox expression (Figure 2.1). 16,20<br />

Two further pieces of evidence demonstrating the anti-inflammatory action of<br />

insulin have emerged recently. First, the treatment of type 2 diabetes with insulin for<br />

2 weeks caused reductions in CRP and MCP-1. 21 Second, the treatment of severe<br />

hyperglycemia associated with marked increases in inflammatory mediators with<br />

insulin resulted in rapid marked decreases in the concentrations of inflammatory<br />

mediators. 22 Most recently, in a rat model in which inflammation was induced with<br />

endotoxin, insulin suppressed the concentration of inflammatory mediators including<br />

Platelet Inhibition<br />

↑ NO release in platelets<br />

↑ e-AMP<br />

Novel Biological<br />

Action of Insulin<br />

Vasodilation Cardio-protective<br />

↑ NO release<br />

↑ eNOS expression<br />

Animals, human<br />

Vascular (other) actions<br />

Anti-apoptotic<br />

Heart, other issues<br />

Anti-oxidant Anti-inflammatory Anti-thrombotic Profibrinolytic Anti-atherosclerotic<br />

↓ ROS generation ↓ NFκB, ↑ lκB<br />

↓ MCP<br />

↓ ICAM-1<br />

↓ CRP<br />

↓ TF ↓ PAI-1 Apo E null mouse<br />

IRS-1 null mouse<br />

IRS-2 null mouse<br />

FIGURE 2.1 Novel biological effects of insulin targeted at endothelial cells, platelets,<br />

and leucocytes resulting in vasodilation, anti-aggregatory effects on platelets, anti-inflammatory<br />

effects, and other related effects. (Source: From Dandona, P. et al., Circulation<br />

111, 1448, 2005.)


The Metabolic Syndrome 19<br />

interleukin-1 beta (IL-1β), IL-6, macrophage inhibitory factor (MIF), and tumor<br />

necrosis factor alpha (TNFα). 23 Insulin also suppressed the expression of pro-inflammatory<br />

transcription factor CEBP and cytokines in the livers of the experimental<br />

animals. Similar reductions in inflammatory mediators were observed in rats with<br />

thermal injuries treated with insulin. 24 Finally, insulin has been shown to suppress<br />

the increases in cytokine concentrations in pigs challenged with endotoxin. 23<br />

The anti-inflammatory, anti-oxidant (ROS-suppressive), anti-thrombotic, and<br />

pro-fibrinolytic effects of insulin have recently been shown to occur in patients with<br />

acute myocardial infarctions when they were treated with low dose infusions of<br />

insulin independently of decreases in glucose concentrations. These patients demonstrated<br />

impressive 40% reductions in plasma CRP and serum amyloid A (SAA)<br />

concentrations at 24 hours. This reduction was maintained at 48 hours of insulin<br />

infusion. 25<br />

These anti-inflammatory effects of insulin have also been shown in patients<br />

undergoing coronary artery bypass grafts in association with extracorporeal circulation.<br />

26 The increase in plasma CRP concentration occurring within 16 hours<br />

of surgery is 30 times greater than that in patients with ST-Elevation Myocardial<br />

Infarction (STEMI). 26 The reduction in the magnitude of increase in CRP and<br />

SAA is 40% — very similar to that observed following insulin infusion in patients<br />

with STEMI. 26<br />

Another novel anti-apoptotic effect of insulin has recently been described. In<br />

experimental acute myocardial infarction in the rat heart, the addition of insulin<br />

to the reperfusion fluid led to a 50% reduction in infarct size. 27 More recently, a<br />

similar cardio-protective effect of insulin was shown in human acute myocardial<br />

infarction when insulin at a low dose was infused with a thrombolytic agent and<br />

heparin. 20 Conversely, insulin-resistant states of obesity and type 2 diabetes have<br />

been shown to be associated with larger infarcts than those observed in nondiabetic<br />

subjects. Further work is required to establish this feature as an integral<br />

component of metabolic syndrome.<br />

It should also be mentioned that insulin administration suppresses atherogenesis<br />

in apolipoprotein E null mice. 28 Conversely, interference with insulin signal<br />

transduction, as in IRS-2 null mice, resulted in atherosclerosis. 29 The IRS-1 null<br />

mouse also has a tendency toward atherosclerosis. It is relevant that a mutation<br />

of IRS-1 (arginine at 792) leads to abnormal vascular reactivity, a decrease in<br />

eNOS expression in endothelial cells, and an increased incidence of coronary<br />

heart disease. 30 It is interesting that the pro-atherogenic effects of insulin proposed<br />

primarily on the basis of in vitro studies are being challenged by evidence<br />

generated in the past 6 years. 7 This debate has been further fueled by two recent<br />

articles showing that knocking out the insulin receptor in myelogenic cells that<br />

are precursors of MNCs (cells that play a crucial role in the pathogenesis of<br />

atherosclerosis) is anti-atherogenic in the background of LDL receptor deficiency<br />

and pro-atherogenic in apo E-deficient animals. 31,32<br />

Consistent with the anti-inflammatory effects of insulin, insulin sensitizers<br />

of the thiazolidinediones class (troglitazone 33,34 and rosiglitazone 35 ) have been<br />

shown to exert anti-inflammatory effects in addition to their glucose lowering


20 Oxidative Stress and Inflammatory Mechanisms<br />

effects in patients with diabetes. Troglitazone has been shown to suppress the<br />

development of diabetes in patients at high risk of developing this condition. 36<br />

Trials are under way to determine whether rosiglitazone and pioglitazone prevent<br />

both type 2 diabetes and atherosclerotic complications. Positive results from those<br />

trials would support the concept that inflammatory mechanisms underlie the<br />

pathogenesis both of insulin resistance and atherosclerosis. It is of interest in this<br />

regard that metformin causes reductions in plasma concentrations of MIF in obese<br />

subjects. 37 Obese individuals have elevated plasma concentrations of this cytokine<br />

and increases in the expression of this cytokine in MNCs. 37 While evidence<br />

indicates that TZDs exert direct anti-inflammatory effects on macrophages in<br />

vitro, it is possible that their effects in vivo may arise through insulin sensitization.<br />

OBESITY <strong>AND</strong> INFLAMMATION<br />

The above data explain why an insulin-resistant state may be pro-inflammatory.<br />

They do not, however, explain the origin of insulin resistance. Mutations of the<br />

genes involved in insulin signal transduction provide one approach to the study<br />

of this issue in humans and in mice with specific gene knockouts. Such lesions<br />

are of interest but are too infrequent to provide a basis for the understanding of<br />

the pathogenesis of insulin resistance at large in humans. Thus, some recent<br />

observations on the interference of insulin signal transduction by inflammatory<br />

mechanisms are of great interest because obesity is a pro-inflammatory state<br />

(Figure 2.2).<br />

Even if we accept that inflammatory mechanisms are involved in the pathogenesis<br />

of interference with insulin signal transduction and of insulin resistance<br />

itself, how does inflammation arise? Over the past decade, obesity has been<br />

associated with inflammation. This association was first proposed in a landmark<br />

paper by Hotamisligil et al. in which TNFα was shown to be constitutively<br />

expressed by adipose tissue, to be hyper-expressed in obesity, and to mediate<br />

insulin resistance in the major animal models of obesity. 38 This seminal paper<br />

also demonstrated that the neutralization of TNFα with soluble TNFα receptors<br />

resulted in the restoration of insulin sensitivity. Thus, the TNFα pro-inflammatory<br />

cytokine was the mediator of insulin resistance. Although the infusion of soluble<br />

TNFα receptors in humans has not reproduced the results observed in mice, 39<br />

Hotamisligil’s paper laid the foundation of the concept that inflammatory mechanisms<br />

may have a role to play in the pathogenesis of insulin resistance.<br />

More data have now accumulated to reinforce the concept that obesity is an<br />

inflammatory state in humans. Increased plasma concentrations of TNFα, IL-6,<br />

CRP, MIF, and other inflammatory mediators were demonstrated in obese subjects.<br />

37,40–44 Adipose tissue has been shown to express most of these pro-inflammatory<br />

mediators. It has also been shown that macrophages residing in adipose<br />

tissue may also be sources of pro-inflammatory factors and may also modulate<br />

the secretory activities of adipocytes. 45


The Metabolic Syndrome 21<br />

Platelet<br />

Hyperaggregability<br />

↓ NO<br />

↓ PGl 2<br />

↑ ROS generation<br />

↑ Oxidative stress<br />

↑ NADPH oxidase<br />

↑ Mitochondrial superoxide<br />

Tonic vasconstriction<br />

-abnormal vascular reactivity<br />

-↓ vascular flow reserve<br />

Chronic pro-inflammatory state<br />

↑ NFκB<br />

↓ IκB<br />

↑ MIF<br />

↑ CRP<br />

↑ TNFα<br />

↓ NO bioavailability<br />

↑ ADMA<br />

Pro-thrombotic state<br />

↓ Cardio-protection<br />

Novel features of metabolic<br />

syndrome on the basis of the<br />

vascular and other effects<br />

of insulin<br />

Larger infarcts<br />

↑ Tendency to CHF<br />

Pro-apoptotic state<br />

↑ Infarct size<br />

Atherosclerosis<br />

CHD<br />

Stroke<br />

Anti-fibrinolytic state<br />

FIGURE 2.2 Extension of metabolic syndrome based on novel actions of insulin. (Source:<br />

From Dandona, P. et al., Circulation 111, 1448, 2005.)<br />

Tissue macrophages are derived from monocytes in blood. Recently, the<br />

mononuclear cells of obese patients, of which monocytes constitute a fraction,<br />

have also been shown to be in an inflammatory state, expressing increased<br />

amounts of pro-inflammatory cytokines and related factors. 46 In addition, these<br />

cells have been shown to have significantly increased binding of NFκB, the key<br />

pro-inflammatory transcription factor, and an increase in the intranuclear expression<br />

of p65 (Rel A), the major protein component of NFκB. These cells also<br />

express diminished amounts of IκBβ, the inhibitor of NFκB. Evidence of inflammation<br />

clearly exists in various cells and in plasma in obesity.<br />

In addition to TNFα and IL-6, the major adipocyte cytokines, three other<br />

important proteins, leptin, adiponectin, and resistin, need mention. While leptin is<br />

known for its function as a satiety signal that inhibits feeding, it has additional roles<br />

as a regulator of sexual function and as an immune modulator. It is also proinflammatory<br />

and induces platelet aggregation. 47–49 Thus, its elevated concentrations<br />

may contribute to the pro-inflammatory state of obesity and to atherogenesis in the<br />

long term. On the other hand, adiponectin, secreted in abundance by adipocytes in<br />

normal subjects, is anti-inflammatory and thus potentially anti-atherogenic. In contrast<br />

to leptin, its concentration falls with weight gain and in obesity. 50,51 It has been<br />

suggested that a low adiponectin concentration may be a marker for atherosclerosis<br />

and coronary heart disease. 52 Furthermore, in several experimental models, it has<br />

been shown to be protective to the arterial endothelium.<br />

↑ TF<br />

↑ PAl-1


22 Oxidative Stress and Inflammatory Mechanisms<br />

Resistin, discovered as a gene suppressed by rosiglitazone in mouse adipose<br />

tissue, earned its name because it induced insulin resistance. 53 Thiazolidenediones<br />

(TZDs) suppress resistin concentrations in humans while inducing an increase in<br />

adiponectin concentration consistent with the anti-inflammatory effects of these<br />

drugs. Furthermore, it has been shown that the increase in adiponectin and the<br />

decrease in resistin occur early after rosiglitazone treatment along with manifestations<br />

of other anti-inflammatory effects prior to any changes in plasma concentrations<br />

of insulin, glucose, or FFAs. Thus, these early anti-inflammatory effects<br />

are independent of the metabolic effects of TZDs. 54<br />

INSULIN RESISTANCE: INFLAMMATORY HYPOTHESIS<br />

Which aspect of the inflammatory state results in insulin resistance? The first of<br />

these potential mechanisms was described by Hotamisligil et al., who demonstrated<br />

that TNFα induced serine phosphorylation of IRS-1 which in turn caused<br />

the serine phosphorylation of the insulin receptor. This prevented the normal<br />

tyrosine phosphorylation of the insulin receptor and thus interfered with insulin<br />

signal transduction. 55 IL-6 and TNFα have recently been shown to induce suppressor<br />

of cytokine signaling-3 (SOCS-3). 56,57 This protein was hitherto thought<br />

to interfere with cytokine signal transduction but is now also known to interfere<br />

with tyrosine phosphorylation of the insulin receptor and insulin receptor substrate-1<br />

(IRS-1) and cause ubiquitination and proteosomal degradation of IRS-<br />

1. 58 This in turn reduces the activation of Akt (protein kinase B) which normally<br />

causes the translocation of the Glut-4 insulin-responsive glucose transporter to<br />

the plasma membrane. It also induces the phosphorylation of the nitric oxide<br />

synthase (NOS) enzyme and its activation to generate NO. 59<br />

A newly described protein known as TRB3 has also been shown to interfere<br />

with the activation of Akt and thus to interfere with insulin action. 60 However, the<br />

association of TRB3 with inflammatory mechanisms has not been demonstrated.<br />

Recent data indicate that Akt2, a key protein involved in insulin signal<br />

transduction that mediates the phosphorylation and activation of e-NOS and NO<br />

secretion, also prevents the mobilization of Rac-1 to cell membranes, thus preventing<br />

superoxide generation. Superoxide generation is dependent upon the<br />

translocation of essential elements of NADPH oxidase (e.g., p47 phox ) from the<br />

cytosol to the membrane. This is mediated by Rac. 61 In the absence of Akt2, there<br />

will be an increase in the translocation of Rac-1 to the membrane, greater formation<br />

of NADPH oxidase complex, increased superoxide generation, and oxidative<br />

stress. It has been shown that Akt2 null mice develop insulin resistance<br />

and mild hyperglycemia in association with hyperinsulinemia. 62<br />

MACRONUTRIENTS <strong>AND</strong> ORIGIN OF INFLAMMATION<br />

If indeed obesity is a pro-inflammatory state and inflammatory mechanisms interfere<br />

with insulin signal transduction, what is the origin of this pro-inflammatory<br />

state? The answer comes mainly from recent observations demonstrating that


The Metabolic Syndrome 23<br />

macronutrient intake may induce oxidative stress and inflammatory responses.<br />

Thus a 75-g glucose challenge has been shown to induce an increase in superoxide<br />

generation by leucocytes by 140% over the basal in addition to increasing p47 phox<br />

expression, a subunit of NADPH oxidase, the enzyme that converts molecular O 2<br />

to superoxide radical. 63<br />

Equicaloric amounts of cream (fat) intake result in similar amounts of oxidative<br />

stress. 64 Glucose intake also results in comprehensive inflammation as<br />

reflected in an increase in intranuclear NFκB binding, a decrease in IκB expression,<br />

and an increase in IKKα and IKKβ, the two kinases that phosphorylate<br />

IκBα and IκBβ and result in their ubiquitination and proteosomal degradation. 65<br />

Glucose intake also causes increases in two other pro-inflammatory transcription<br />

factors: AP-1 and Egr-1. 66 AP-1 regulates the transcription of matrix metalloproteinases<br />

while Egr-1 modulates the transcription of tissue factor (TF) and PAI-1.<br />

Thus, glucose intake increases the expression of matrix metalloproteinases 2 and<br />

9 as well as expression of TF and PAI-1.<br />

A mixed meal from a fast food chain was also shown to induce the activation<br />

of NFκB, a reduction in IκBα, an increase in IKKα and IKKβ, and an increase<br />

in superoxide radical generation by MNCs. 67 It is also of interest that an intravenous<br />

infusion of triglyceride with heparin in normal subjects with elevations<br />

of FFA concentrations to levels comparable to those found in obese subjects<br />

resulted in inflammatory responses. 14<br />

All genes that are stimulated by acute nutritional intake have also been shown<br />

to be activated in the basal states of obese subjects such that the concentrations<br />

of these gene products were elevated. Consistent with this, reductions in macronutrient<br />

intake in obese subjects (1000 kcal daily for 4 weeks) were shown to<br />

reduce both oxidative stress and inflammatory mediators. 8 Similarly, a 48-hour<br />

fast has been shown to reduce ROS generation by over 50% in normal subjects;<br />

the expression of p47 phox was also reduced. 68<br />

Clearly, macronutrient intake is a major regulator of oxidative stress. It is<br />

relevant that the superoxide radical generated during oxidative stress is an activator<br />

of at least two major pro-inflammatory transcription factors, NFκB and AP-<br />

1. NFκB regulates the transcriptional activities of at least 200 genes, most of<br />

which are pro-inflammatory. 69–72 Thus, it is not surprising that obesity is a proinflammatory<br />

condition. Indeed, the MNCs of obese individuals are in a proinflammatory<br />

state, expressing an excess of a series of pro-inflammatory genes<br />

in addition to demonstrating increased NFκB binding, p65 expression, and<br />

decreased IκBβ protein. 46<br />

In addition to obesity and increased macronutrient intake, genetic and other<br />

environmental factors may induce the activation of inflammatory mechanisms and<br />

the induction of oxidative stress. These factors may be relevant in those ethnic<br />

groups in whom metabolic syndrome has been shown to occur in the absence of<br />

obesity. In these groups, migration to western countries like the U.S. and U.K.<br />

results in increased adiposity with a sedentary lifestyle that results in a phenotype<br />

of the metabolic syndrome against an appropriate genetic background (Figure 2.3).


24 Oxidative Stress and Inflammatory Mechanisms<br />

Obesity<br />

↑ Food Intake<br />

↑ ROS, ↑ p47<br />

(Oxidative Stress and Inflammation)<br />

phos<br />

↑ NFκB<br />

↑ AP-1, ↑ MMPs<br />

↑ Egr-1, ↑ TF, ↑ PAI-1<br />

Serine Phosphorylation of IRS-1<br />

↑ SOCS-3<br />

Interruption of Insulin Signal Transduction<br />

Insulin Resistance<br />

↑ Lipolysis<br />

↑ FFA<br />

Genetic Factors<br />

FIGURE 2.3 Pathogenesis of metabolic syndrome. (Source: From Dandona, P. et al.,<br />

Circulation 111, 1448, 2005.)<br />

When considering macronutrient-induced inflammation, it can be argued that<br />

the foods consumed now were always consumed, so why are their pro-inflammatory<br />

effects suddenly becoming relevant? The reason is that the amounts of food consumed<br />

now are far greater; furthermore, larger portions of the average diet consist<br />

of fast foods and lack sufficient fiber, fruits, and vegetables. Furthermore, the<br />

relative contents of refined carbohydrates and saturated fats have increased. This<br />

combination results in the inability of endogenously secreted insulin in response<br />

to meal intake to suppress the inflammation generated by the meal.<br />

It is of interest in this regard that a 900-kcal AHA step 2 diet-based meal<br />

rich in fruit and fiber does not cause significant oxidative stress or inflammation<br />

in contrast to the effect of an isocaloric fast food meal. 73 Furthermore, orange<br />

juice taken in amounts equivalent to 75 g glucose (= 300 cal) does not induce<br />

any increase in O 2 • generation or an increase in NFB binding by MNCs. This<br />

may be due partly to the anti-oxidant and anti-inflammatory effects of ascorbic<br />

acid and flavanoids contained in orange juice. However, it has also been shown<br />

that fructose, which accounts for 50% of carbohydrates in orange juice, does not<br />

cause O 2 • generation or increased NFB binding. It is also of interest that orange<br />

juice induced a greater increase in insulin for a certain increase in glucose<br />

concentration when compared to glucose challenge. 74 This raises a novel concept<br />

when considering post-prandial hyperglycemia, oxidative stress, and inflammation<br />

following ingestion of various foods. The recent demonstration that M3<br />

muscarinic receptors in the β cells of the pancreatic islets enhanced the insulinogenic<br />

response to a given glucose load or concentration suggests such an avenue


The Metabolic Syndrome 25<br />

of investigation. It is possible that certain foods like orange juice activate and/or<br />

increase the expression of M3 muscarinic receptors on β cells to enhance insulin<br />

secretion in response to glucose challenge. 75<br />

Both glucose and cream intake induce the expression of a series of proinflammatory<br />

genes in peripheral blood MNCs as reflected in the mRNA. These<br />

genes include TNFα, IL-6, and other related genes. Among the genes induced<br />

are SOCS-3, the molecule considered a potential candidate for the mediation of<br />

insulin resistance through serine phosphorylation, ubiquitination, and proteasomal<br />

degradation of IRS-1.<br />

The increase in superoxide radical generation also results in diminished<br />

bioavailability of NO since NO binds to superoxide radical to form peroxynitrate.<br />

76 In addition to the fact that Akt is inhibited due to insulin resistance, and<br />

thus NOS is also inhibited, the reduction in NO bioavailability can result in a<br />

marked reduction in NO action. Furthermore, TNFα suppresses the expression<br />

of NOS. The factors result in abnormalities in endothelium-mediated vasodilatation<br />

and vascular reactivity. 77 Interestingly, the abnormalities in vascular reactivity<br />

in the obese insulin-resistant population can be reproduced acutely by a 900-kcal<br />

fast food meal, just as the pro-inflammatory changes in obesity can be reproduced<br />

by a similar meal. 67 It is noteworthy that the plasma concentrations of asymmetric<br />

dimethylarginine (ADMA) are elevated in obesity and inhibit NOS activity, thus<br />

reducing the synthesis and secretion of NO. 78 Rosiglitazone suppresses plasma<br />

ADMA concentrations while improving the impaired vascular reactivity in the<br />

obese and those with type 2 diabetes. 79<br />

While the initial work on macronutrient intake with glucose, cream, and fast<br />

food meals showed a pro-inflammatory effect associated with oxidative stress,<br />

data are now emerging to demonstrate that some macronutrients may be safe and<br />

non-inflammatory. Thus, a 900-kcal breakfast rich in fruit and fiber does not cause<br />

oxidative stress or inflammation. The intake of vitamin E prior to glucose challenge<br />

also suppresses oxidative stress and inflammation. Similarly, alcohol 65 and<br />

orange juice in equicaloric amounts do not cause oxidative stress or inflammation.<br />

Because orange juice is rich in flavanoids and vitamin C, it is possible that the<br />

presence of macronutrients in food may alter or suppress oxidative stress or<br />

inflammation. Furthermore, data show that vitamin E administration to patients<br />

with insulin resistance reduced cytokine production by MNCs. 80<br />

Other strategies to prevent post-prandial oxidative stress and inflammation<br />

may involve drug therapies. One initial attempt in this area indicates that rosiglitazone<br />

treatment for 6 weeks (8 mg/day) led to the total suppression of oxidative<br />

and inflammatory stress caused by a 75-g glucose challenge in type 2 diabetes. 81<br />

CONCLUSION: INFLAMMATION HYPOTHESIS OF<br />

METABOLIC SYNDROME<br />

In conclusion, the pro-inflammatory state of obesity and metabolic syndrome<br />

originates with excessive caloric intake and is probably due to over-nutrition in


26 Oxidative Stress and Inflammatory Mechanisms<br />

a majority of patients in the U.S. The pro-inflammatory state induces insulin<br />

resistance leading to clinical and biochemical manifestations of the metabolic<br />

syndrome. This resistance to insulin action further promotes inflammation<br />

through increases in lipolysis and plasma FFA concentrations on one hand and<br />

interference with the anti-inflammatory effect of insulin on the other.<br />

While these factors may be the most important ones in a majority of patients<br />

with metabolic syndrome, it is possible that genetic factors may also contribute<br />

to the inflammatory stress in metabolic syndrome. These factors may be important<br />

in ethnic groups such as Asian Indians who may have increased amounts of upper<br />

abdominal fat in spite of normal BMI values. 82 Since excessive nutritional intake<br />

probably accounts for the inflammation at least in obesity-associated metabolic<br />

syndrome, the most rational way to suppress such inflammation is through caloric<br />

restriction.<br />

The other lifestyle change that affects inflammation is exercise. Exercise<br />

produces a decrease in the indices of inflammation such as plasma CRP concentration.<br />

83 The mechanism underlying this effect is not known. However, it is<br />

noteworthy that a lifestyle change is a very effective way to reduce the rate of<br />

development of diabetes in a pre-diabetic population as shown by the diabetes<br />

prevention study. 84,85 Exercise and reductions in macronutrient intake both cause<br />

a reduction in inflammation. Several drugs have also been shown to reduce<br />

oxidative stress and inflammation and may therefore be used in the treatment of<br />

metabolic syndrome. Among them are thiazolidinediones, angiotensin II receptor<br />

blockers, and carvedilol, a β-blocker with some α-blocking activity. 33,35,54,86–88<br />

REFERENCES<br />

1. Reaven GM. Banting Lecture 1988: role of insulin resistance in human disease.<br />

Diabetes 37, 1595, 1988.<br />

2. Pyorala K. Relationship of glucose tolerance and plasma insulin to the incidence<br />

of coronary heart disease: results from two population studies in Finland. Diabetes<br />

Care 2, 131, 1979.<br />

3. Ninomiya JK, L’Italien G, Criqui MH et al. Association of the metabolic syndrome<br />

with history of myocardial infarction and stroke in the third national health and<br />

nutrition examination survey. Circulation 109, 42, 2004.<br />

4. Lakka HM, Laaksonen DE, Lakka TA et al. The metabolic syndrome and total<br />

and cardiovascular disease mortality in middle-aged men. JAMA 288, 2709, 2002.<br />

5. Festa A, D’Agostino R, Jr, Howard G et al. Chronic subclinical inflammation as<br />

part of the insulin resistance syndrome: the Insulin Resistance Atherosclerosis<br />

Study (IRAS). Circulation 102, 42, 2000.<br />

6. Klein BE, Klein R, and Lee KE. Components of the metabolic syndrome and risk<br />

of cardiovascular disease and diabetes in beaver dam. Diabetes Care 25, 1790,<br />

2002.<br />

7. Dandona P, Aljada A, and Mohanty P. The anti-inflammatory and potential antiatherogenic<br />

effect of insulin: a new paradigm. Diabetologia 45, 924, 2002.


The Metabolic Syndrome 27<br />

8. Dandona P, Mohanty P, Ghanim H et al. The suppressive effect of dietary restriction<br />

and weight loss in the obese on the generation of reactive oxygen species by<br />

leukocytes, lipid peroxidation, and protein carbonylation. J Clin Endocrinol Metab<br />

86, 355, 2001.<br />

9. Lin KY, Ito A, Asagami T et al. Impaired nitric oxide synthase pathway in diabetes<br />

mellitus: role of asymmetric dimethylarginine and dimethylarginine dimethylaminohydrolase.<br />

Circulation 106, 987, 2002.<br />

10. Esler M, Rumantir M, Kaye D et al. The sympathetic neurobiology of essential<br />

hypertension: disparate influences of obesity, stress, and noradrenaline transporter<br />

dysfunction? Am J Hypertens 14, 139S, 2001.<br />

11. Giacchetti G, Faloia E, Sardu C et al. Gene expression of angiotensinogen in<br />

adipose tissue of obese patients. Int J Obes Relat Metab Disord 24, Suppl 2, S142,<br />

2000.<br />

12. Groenendijk M, Cantor RM, Blom NH et al. Association of plasma lipids and<br />

apolipoproteins with the insulin response element in the apoC-III promoter region<br />

in familial combined hyperlipidemia. J Lipid Res 40, 1036, 1999.<br />

13. Mooradian AD, Haas MJ, and Wong NC. Transcriptional control of apolipoprotein<br />

A-I gene expression in diabetes. Diabetes 53, 513, 2004.<br />

14. Tripathy D, Mohanty P, Dhindsa S et al. Elevation of free fatty acids induces<br />

inflammation and impairs vascular reactivity in healthy subjects. Diabetes 52,<br />

2882, 2003.<br />

15. Dandona P, Aljada A, and Bandyopadhyay A. Inflammation: the link between<br />

insulin resistance, obesity and diabetes. Trends Immunol 25, 4, 2004.<br />

16. Dandona P, Aljada A, Mohanty P et al. Insulin inhibits intranuclear nuclear factor<br />

kappa-B and stimulates I-kappa-B in mononuclear cells in obese subjects: evidence<br />

for an anti-inflammatory effect? J Clin Endocrinol Metab 86, 3257, 2001.<br />

17. Aljada A, Ghanim H, Mohanty P et al. Insulin inhibits the pro-inflammatory<br />

transcription factor early growth response gene-1 (Egr)-1 expression in mononuclear<br />

cells (MNC) and reduces plasma tissue factor (TF) and plasminogen activator<br />

inhibitor-1 (PAI-1) concentrations. J Clin Endocrinol Metab 87, 1419, 2002.<br />

18. Ghanim H, Mohanty P, Aljada A et al. Insulin reduces the pro-inflammatory<br />

transcription factor, activation protein-1 (AP-1), in mononuclear cells (MNC) and<br />

plasma matrix metalloproteinase-9 (MMP-9) concentration. Diabetes 50, Suppl<br />

2, A408, 2001.<br />

19. Dandona P, Aljada A, Mohanty P et al. Insulin suppresses plasma concentration<br />

of vascular endothelial growth factor and matrix metalloproteinase-9. Diabetes<br />

Care 26, 3310, 2003.<br />

20. Chaudhuri A, Janicke D, Wilson MF et al. Anti-inflammatory and profibrinolytic<br />

effect of insulin in acute ST-segment-elevation myocardial infarction. Circulation<br />

109, 849, 2004.<br />

21. Takebayashi K, Aso Y, and Inukai T. Initiation of insulin therapy reduces serum<br />

concentrations of high-sensitivity C-reactive protein in patients with type 2 diabetes.<br />

Metabolism 53, 693, 2004.<br />

22. Stentz FB, Umpierrez GE, Cuervo R et al. Proinflammatory cytokines, markers<br />

of cardiovascular risks, oxidative stress, and lipid peroxidation in patients with<br />

hyperglycemic crises. Diabetes 53, 2079, 2004.<br />

23. Jeschke MG, Klein D, Bolder U et al. Insulin attenuates the systemic inflammatory<br />

response in endotoxemic rats. Endocrinology 145, 4084, 2004.


28 Oxidative Stress and Inflammatory Mechanisms<br />

24. Jeschke MG, Einspanier R, Klein D et al. Insulin attenuates the systemic inflammatory<br />

response to thermal trauma. Mol Med. 8, 443, 2002.<br />

25. Chaudhuri A, Janicke D, Wilson MF et al. Anti-inflammatory and pro-fibrinolytic<br />

effect of insulin in acute ST-elevation myocardial infarction. Circulation 109, 849,<br />

2004.<br />

26. Visser L, Zuurbier CJ, Hoek FJ et al. Glucose, insulin and potassium applied as<br />

perioperative hyperinsulinaemic normoglycaemic clamp: effects on inflammatory<br />

response during coronary artery surgery. Br J Anaesth 95, 448, 2005.<br />

27. Jonassen AK, Sack MN, Mjos OD et al. Myocardial protection by insulin at<br />

reperfusion requires early administration and is mediated via Akt and p70s6 kinase<br />

cell-survival signaling. Circ Res 89, 1191, 2001.<br />

28. Shamir R, Shehadeh N, Rosenblat M et al. Oral insulin supplementation attenuates<br />

atherosclerosis progression in apolipoprotein E-deficient mice. Arterioscler<br />

Thromb Vasc Biol 23, 104, 2003.<br />

29. Kubota T, Kubota N, Moroi M et al. Lack of insulin receptor substrate-2 causes<br />

progressive neointima formation in response to vessel injury. Circulation 107,<br />

3073, 2003.<br />

30. Federici M, Pandolfi A, De Filippis EA et al. G972R IRS-1 variant impairs insulin<br />

regulation of endothelial nitric oxide synthase in cultured human endothelial cells.<br />

Circulation 109, 399, 2004.<br />

31. Han S, Liang CP, DeVries-Seimon T et al. Macrophage insulin receptor deficiency<br />

increases ER stress-induced apoptosis and necrotic core formation in advanced<br />

atherosclerotic lesions. Cell Metab 3, 257, 2006.<br />

32. Baumgartl J, Baudler S, Scherner M et al. Myeloid lineage cell-restricted insulin<br />

resistance protects apolipoprotein E-deficient mice against atherosclerosis. Cell<br />

Metab 3, 247, 2006.<br />

33. Ghanim H, Garg R, Aljada A et al. Suppression of nuclear factor kappa-B and<br />

stimulation of inhibitor kappa-B by troglitazone: evidence for an anti-inflammatory<br />

effect and a potential anti-atherosclerotic effect in the obese. J Clin Endocrinol<br />

Metab 86, 1306, 2001.<br />

34. Aljada A, Garg R, Ghanim H et al. Nuclear factor-kappa-B suppressive and<br />

inhibitor-kappa-B stimulatory effects of troglitazone in obese patients with type<br />

2 diabetes: evidence of an anti-inflammatory action? J Clin Endocrinol Metab 86,<br />

3250, 2001.<br />

35. Mohanty P, Aljada A, Ghanim H et al. Evidence for a potent antiinflammatory<br />

effect of rosiglitazone. J Clin Endocrinol Metab 89, 2728, 2004.<br />

36. Buchanan TA, Xiang AH, Peters RK et al. Preservation of pancreatic beta-cell<br />

function and prevention of type 2 diabetes by pharmacological treatment of insulin<br />

resistance in high-risk Hispanic women. Diabetes 51, 2796, 2002.<br />

37. Dandona P, Aljada A, Ghanim H et al. Increased plasma concentration of macrophage<br />

migration inhibitory factor (MIF) and MIF mRNA in mononuclear cells in<br />

the obese and the suppressive action of metformin. J Clin Endocrinol Metab 89,<br />

5043, 2004.<br />

38. Hotamisligil GS, Shargill NS, and Spiegelman BM. Adipose expression of tumor<br />

necrosis factor-alpha: direct role in obesity-linked insulin resistance. Science 259,<br />

87, 1993.<br />

39. Ofei F, Hurel S, Newkirk J et al. Effects of an engineered human anti-TNF-alpha<br />

antibody (CDP571) on insulin sensitivity and glycemic control in patients with<br />

NIDDM. Diabetes 45, 881, 1996.


The Metabolic Syndrome 29<br />

40. Kern PA, Ranganathan S, Li C et al. Adipose tissue tumor necrosis factor and<br />

interleukin-6 expression in human obesity and insulin resistance. Am J Physiol<br />

Endocrinol Metab 280, E745, 2001.<br />

41. Dandona P, Weinstock R, Thusu K et al. Tumor necrosis factor-alpha in sera of<br />

obese patients: fall with weight loss. J Clin Endocrinol Metab 83, 2907, 1998.<br />

42. Vozarova B, Weyer C, Hanson K et al. Circulating interleukin-6 in relation to<br />

adiposity, insulin action, and insulin secretion. Obes Res 9, 414, 2001.<br />

43. Wakabayashi I. Age-related change in relationship between body-mass index,<br />

serum sialic acid, and atherogenic risk factors. J Atheroscler Thromb 5, 60, 1998.<br />

44. Pradhan AD, Manson JE, Rifai N et al. C-reactive protein, interleukin-6, and risk<br />

of developing type 2 diabetes mellitus. JAMA 286, 327, 2001.<br />

45. Xu H, Barnes GT, Yang Q et al. Chronic inflammation in fat plays a crucial role<br />

in the development of obesity-related insulin resistance. J Clin Invest 112, 1821,<br />

2003.<br />

46. Ghanim H, Aljada A, Hofmeyer D et al. The circulating mononuclear cells in the<br />

obese are in a pro-inflammatory state. Circulation 110, 1564, 2004.<br />

47. La Cava A, Alviggi C, and Matarese G. Unraveling the multiple roles of leptin in<br />

inflammation and autoimmunity. J Mol Med 82, 4, 2004.<br />

48. Huang L and Li C. Leptin: a multifunctional hormone. Cell Res. 10, 81, 2000.<br />

49. Nakata M, Yada T, Soejima N et al. Leptin promotes aggregation of human<br />

platelets via the long form of its receptor. Diabetes 48, 426, 1999.<br />

50. Kubota N, Terauchi Y, Yamauchi T et al. Disruption of adiponectin causes insulin<br />

resistance and neointimal formation. J Biol Chem 277, 25863, 2002.<br />

51. Ukkola O and Santaniemi M. Adiponectin: a link between excess adiposity and<br />

associated comorbidities? J Mol Med 80, 696, 2002.<br />

52. Pischon T, Girman CJ, Hotamisligil GS et al. Plasma adiponectin levels and risk<br />

of myocardial infarction in men. JAMA 291, 1730, 2004.<br />

53. Steppan CM, Bailey ST, Bhat S et al. The hormone resistin links obesity to<br />

diabetes. Nature 409, 307, 2001.<br />

54. Ghanim H, Dhindsa S, Aljada A et al. Low dose rosiglitazone exerts an antiinflammatory<br />

effect with an increase in adiponectin independently of free fatty<br />

acid (FFA) fall and insulin sensitization in obese type 2 diabetics. J Clin Endocrinol<br />

Metab June 27, 2006.<br />

55. Hotamisligil GS, Peraldi P, Budavari A et al. IRS-1-mediated inhibition of insulin<br />

receptor tyrosine kinase activity in TNF-alpha- and obesity-induced insulin resistance.<br />

Science 271, 665, 1996.<br />

56. Senn JJ, Klover PJ, Nowak IA et al. Suppressor of cytokine signaling-3 (SOCS-<br />

3), a potential mediator of interleukin-6-dependent insulin resistance in hepatocytes.<br />

J Biol Chem 278, 13740, 2003.<br />

57. Emanuelli B, Peraldi P, Filloux C et al. SOCS-3 inhibits insulin signaling and is<br />

up-regulated in response to tumor necrosis factor-alpha in the adipose tissue of<br />

obese mice. J Biol Chem 276, 47944, 2001.<br />

58. Rui L, Yuan M, Frantz D et al. SOCS-1 and SOCS-3 block insulin signaling by<br />

ubiquitin-mediated degradation of IRS1 and IRS2. J Biol Chem 277, 42394, 2002.<br />

59. Dimmeler S, Fleming I, Fisslthaler B et al. Activation of nitric oxide synthase in<br />

endothelial cells by Akt-dependent phosphorylation. Nature 399, 601, 1999.<br />

60. Du K, Herzig S, Kulkarni RN et al. TRB3: a tribbles homolog that inhibits<br />

Akt/PKB activation by insulin in liver. Science 300, 1574, 2003.


30 Oxidative Stress and Inflammatory Mechanisms<br />

61. Ozaki M, Haga S, Zhang HQ et al. Inhibition of hypoxia/reoxygenation-induced<br />

oxidative stress in HGF-stimulated antiapoptotic signaling: role of PI3-K and Akt<br />

kinase upon rac1. Cell Death Differ 10, 508, 2003.<br />

62. Cho H, Mu J, Kim JK et al. Insulin resistance and a diabetes mellitus-like<br />

syndrome in mice lacking the protein kinase Akt2 (PKB beta). Science 292, 1728,<br />

2001.<br />

63. Mohanty P, Hamouda W, Garg R et al. Glucose challenge stimulates reactive<br />

oxygen species (ROS) generation by leucocytes. J Clin Endocrinol Metab 85,<br />

2970, 2000.<br />

64. Mohanty P, Ghanim H, Hamouda W et al. Both lipid and protein intakes stimulate<br />

increased generation of reactive oxygen species by polymorphonuclear leukocytes<br />

and mononuclear cells. Am J Clin Nutr 75, 767, 2002.<br />

65. Dhindsa S, Tripathy D, Mohanty P et al. Differential effects of glucose and alcohol<br />

on reactive oxygen species generation and intranuclear nuclear factor-kappa-B in<br />

mononuclear cells. Metabolism 53, 330, 2004.<br />

66. Aljada A, Ghanim H, Mohanty P et al. Glucose intake induces an increase in AP-<br />

1 and Egr-1 binding activities and tissue factor and matrix metalloproteinase<br />

expressions in mononuclear cells and plasma tissue factor and matrix metalloproteinase<br />

concentrations. Am J Clin Nutr 80, 51, 2004.<br />

67. Aljada A, Mohanty P, Ghanim H et al. Increase in intranuclear nuclear factor<br />

kappaB and decrease in inhibitor kappaB in mononuclear cells after a mixed meal:<br />

evidence for a proinflammatory effect. Am J Clin Nutr 79, 682, 2004.<br />

68. Dandona P, Mohanty P, Hamouda W et al. Inhibitory effect of a two-day fast on<br />

reactive oxygen species (ROS) generation by leucocytes and plasma ortho-tyrosine<br />

and meta-tyrosine concentrations. J Clin Endocrinol Metab 86, 2899, 2001.<br />

69. Woronicz JD, Gao X, Cao Z et al. I-kappa-B kinase-beta: NF-kappa-B activation<br />

and complex formation with I-kappa-B kinase-alpha and NIK. Science 278, 866,<br />

1997.<br />

70. Wang S, Leonard SS, Castranova V et al. The role of superoxide radical in TNFalpha<br />

induced NF-kappa-B activation. Ann Clin Lab Sci 29, 192, 1999.<br />

71. Verma IM, Stevenson JK, Schwarz EM et al. Rel/NF-kappa B/I-kappa B family:<br />

intimate tales of association and dissociation. Genes Dev 9, 2723, 1995.<br />

72. Velasco M, Diaz-Guerra MJ, Martin-Sanz P et al. Rapid up-regulation of I-kappa-<br />

B and abrogation of NF-kappa-B activity in peritoneal macrophages stimulated<br />

with lipopolysaccharide. J Biol Chem 272, 23025, 1997.<br />

73. Mohanty P, Daoud N, Ghanim H et al. Absence of oxidative stress and inflammation<br />

following the intake of a 900 kcal meal rich in fruit and fiber. Diabetes<br />

53, Suppl 2, A405, 2004.<br />

74. Ghanim H, Mohanty P, Pathak R et al. Caloric intake from orange juice does not<br />

induce oxidative and inflammatory response. Boston, Endocrine Society, 88th<br />

Annual Meeting, 2006.<br />

75. Gautam D, Han SJ, Hamdan FF et al. A critical role for beta cell M3 muscarinic<br />

acetylcholine receptors in regulating insulin release and blood glucose homeostasis<br />

in vivo. Cell Metab 3, 449, 2006.<br />

76. Koppenol WH, Moreno JJ, Pryor WA et al. Peroxynitrite, a cloaked oxidant formed<br />

by nitric oxide and superoxide. Chem Res Toxicol 5, 834, 1992.<br />

77. Fard A, Tuck CH, Donis JA et al. Acute elevations of plasma asymmetric dimethylarginine<br />

and impaired endothelial function in response to a high-fat meal in<br />

patients with type 2 diabetes. Arterioscler Thromb Vasc Biol 20, 2039, 2000.


The Metabolic Syndrome 31<br />

78. Chan NN and Chan JC. Asymmetric dimethylarginine (ADMA): a potential link<br />

between endothelial dysfunction and cardiovascular diseases in insulin resistance<br />

syndrome? Diabetologia 45, 1609, 2002.<br />

79. Stuhlinger MC, Abbasi F, Chu JW et al. Relationship between insulin resistance<br />

and an endogenous nitric oxide synthase inhibitor. JAMA 287, 1420, 2002.<br />

80. Devaraj S and Jialal I. Low-density lipoprotein postsecretory modification, monocyte<br />

function, and circulating adhesion molecules in type 2 diabetic patients with<br />

and without macrovascular complications: the effect of alpha-tocopherol supplementation.<br />

Circulation 102, 191, 2000.<br />

81. Ahmad-Aljada SD, Ghanim H, Garg R et al. Rosiglitazone treatment reduces<br />

glucose-induced oxidative stress and inflammation. Presented at Endocrine Society<br />

87th Annual Meeting, San Diego, 2005.<br />

82. Banerji MA, Faridi N, Atluri R et al. Body composition, visceral fat, leptin, and<br />

insulin resistance in Asian Indian men. J Clin Endocrinol Metab 84, 137, 1999.<br />

83. Church TS, Barlow CE, Earnest CP et al. Associations between cardiorespiratory<br />

fitness and C-reactive protein in men. Arterioscler Thromb Vasc Biol 22, 1869,<br />

2002.<br />

84. Tuomilehto J, Lindstrom J, Eriksson JG et al. Prevention of type 2 diabetes mellitus<br />

by changes in lifestyle among subjects with impaired glucose tolerance. New Engl<br />

J Med 344, 1343, 2001.<br />

85. Knowler WC, Barrett-Connor E, Fowler SE et al. Reduction in the incidence of<br />

type 2 diabetes with lifestyle intervention or metformin. New Engl J Med 346,<br />

393, 2002.<br />

86. Dandona P, Karne R, Ghanim H et al. Carvedilol inhibits reactive oxygen species<br />

generation by leukocytes and oxidative damage to amino acids. Circulation 101,<br />

122, 2000.<br />

87. Garg R, Kumbkarni Y, Aljada A et al. Troglitazone reduces reactive oxygen species<br />

generation by leukocytes and lipid peroxidation and improves flow-mediated<br />

vasodilatation in obese subjects. Hypertension 36, 430, 2000.<br />

88. Dandona P, Kumar V, Aljada A et al. Angiotensin II receptor blocker valsartan<br />

suppresses reactive oxygen species generation in leukocytes, nuclear factor-kappa<br />

B, in mononuclear cells of normal subjects: evidence of an antiinflammatory<br />

action. J Clin Endocrinol Metab 88, 4496, 2003.


3<br />

CONTENTS<br />

The Role of Oxidative<br />

Stress in Diseases<br />

Associated with<br />

Overweight and<br />

Obesity<br />

Ginger L. Milne, Ling Gao, Joshua D. Brooks,<br />

and Jason D. Morrow<br />

Introduction .........................................................................................................33<br />

F 2-Isoprostanes as Markers of Oxidant Stress In Vivo.......................................34<br />

F 2-Isoprostanes and Overweight and Obesity ....................................................35<br />

Therapeutic Targets for Reducing Oxidant Stress in Overweight and<br />

Obese Patients ...........................................................................................38<br />

Antioxidants...............................................................................................38<br />

ω-3 Polyunsaturated Fatty Acids ..............................................................39<br />

Conclusions .........................................................................................................41<br />

Acknowledgments ...............................................................................................42<br />

References ...........................................................................................................42<br />

INTRODUCTION<br />

The marked increase in the incidence of overweight and obese persons is recognized<br />

as perhaps the most serious public health issue in the United States. It is<br />

estimated that two-thirds of American adults are overweight and nearly 30% are<br />

obese. 1,2 Additionally, the incidence of overweight and obesity in children and<br />

adolescents is rising; in 2004, it was estimated that 16% of youth are either<br />

overweight or obese. 1,3 Both morbidity and mortality increase with excessive<br />

body weight. 4–6<br />

33


34 Oxidative Stress and Inflammatory Mechanisms<br />

Multiple studies have shown that the risks of developing cardiovascular disease,<br />

type 2 diabetes mellitus, hypertension, dyslipidemia, gallbladder disease,<br />

osteoarthritis, stroke, and certain types of cancers increase with degree of overweight<br />

in both men and women. 2,7 Furthermore, early onset of symptoms of these<br />

diseases that were once only associated with adulthood are now observed in<br />

overweight adolescents and include high blood pressure, atherosclerosis, and type<br />

2 diabetes mellitus. 8 Despite the well characterized association of overweight and<br />

disease incidence, the mechanisms by which overweight contributes to disease<br />

pathology are poorly understood. Nonetheless, several reports provide evidence<br />

that elevated systemic oxidant stress may be an important mechanism by which<br />

obesity promotes the development of chronic human disease. 9–11<br />

This chapter will examine the measurement of F 2-isoprostanes (IsoPs), a<br />

class of oxidized lipids generated from the peroxidation of arachidonic acid, as<br />

a marker of oxidant stress in overweight, obesity, and associated diseases. It<br />

will also assess the impacts of select therapeutic interventions on F 2-IsoP formation<br />

in this population.<br />

F 2-ISOPROSTANES AS MARKERS OF OXIDANT <strong>STRESS</strong> IN VIVO<br />

F 2-IsoPs represent one class of oxidized lipids formed in vivo in humans. 12–14<br />

These compounds that were first identified by our laboratory in 1990 are prostaglandin<br />

(PG)-like molecules generated non-enzymatically from the free radicalinitiated<br />

peroxidation of arachidonic acid, a ubiquitous polyunsaturated fatty<br />

acid. 15 Figure 3.1 shows the mechanism of formation of the F 2-IsoPs. Briefly,<br />

after abstraction of a bis-allylic hydrogen atom, one molecule of oxygen adds to<br />

arachidonic acid to form a peroxyl radical. This peroxyl radical then undergoes<br />

5-exo cyclization and subsequent addition of another molecule of oxygen to form<br />

PGG 2-like compounds.<br />

The unstable bicyclic endoperoxide intermediates are then reduced to the F 2-<br />

IsoPs, isomers of the cyclooxygenase-derived product PGF 2α. Based on this mechanism<br />

of formation, four series of F 2-IsoP regioisomers are generated. In total, 64<br />

different F 2-IsoP stereoisomers are formed from the oxidation of arachidonic acid.<br />

In more recent years since their discovery, several methods have been developed<br />

to quantify the formation of F 2-IsoPs in vivo. 16 The method our laboratory<br />

and others found to be most sensitive, specific, and reliable employs the use of<br />

mass spectrometry and stable isotope dilution techniques. 17 Employing this technique,<br />

F 2-IsoPs have been detected in all human tissues and fluids examined,<br />

which is of particular importance in that it allows for an assessment of the effects<br />

of diseases on endogenous oxidant tone. 14 Additionally, defining levels of F 2-<br />

IsoPs in vivo is important because it allows for the determination of the extent<br />

to which various therapeutic interventions affect levels of oxidant stress.<br />

For human studies, the quantification of F 2-IsoPs in body fluids such as urine<br />

and plasma is significantly more convenient and less invasive than measuring<br />

these compounds in organ tissue. Based on available data, quantification of these<br />

compounds in either plasma or urine is representative of their endogenous


Oxidative Stress in Diseases Associated with Overweight and Obesity 35<br />

O O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

HO<br />

10<br />

HO<br />

COOH<br />

COOH<br />

COOH<br />

O<br />

O<br />

Arachidonic acid<br />

O<br />

OH<br />

COOH<br />

O<br />

O<br />

O O<br />

COOH COOH<br />

COOH<br />

O<br />

O<br />

FIGURE 3.1 Mechanism of F 2-isoprostane formation from the free radical-initiated peroxidation<br />

of arachidonic acid.<br />

production, and thus gives a highly precise and accurate index of in vivo oxidant<br />

stress. 14,16,18,19 In fact, since their initial discovery, quantification of the F 2-IsoPs<br />

by mass spectrometry has emerged as the “gold standard” index of in vivo oxidant<br />

stress status.<br />

Recently, this assertion was independently confirmed by a National Institutes<br />

of Health-sponsored program termed the Biomarkers of Oxidative Stress Study<br />

(BOSS) that directly compared measurements of plasma and urinary F 2-IsoPs<br />

with other well known but less robust biomarkers of oxidant stress, including<br />

malondialdehyde (MDA) and other measures of lipid peroxidation, plasma glutathione,<br />

plasma antioxidant levels, protein carbonyls, and 8-hydroxy-deoxyguanosine.<br />

20,21<br />

F 2-ISOPROSTANES <strong>AND</strong> OVERWEIGHT <strong>AND</strong> OBESITY<br />

As discussed above, measurement of plasma or urinary F 2-IsoPs allows for an<br />

assessment of the effects of diseases on oxidant tone in vivo. In the past decade,<br />

O O<br />

COOH O<br />

O<br />

O2 O2 O2 O2 OO<br />

COOH<br />

O<br />

O<br />

COOH<br />

O<br />

OO<br />

O<br />

COOH<br />

O<br />

OO<br />

O<br />

OO<br />

OOH<br />

COOH<br />

O<br />

O<br />

COOH<br />

O<br />

OOH<br />

O<br />

COOH<br />

O<br />

OOH O<br />

OOH<br />

[H] [H] [H] [H]<br />

COOH<br />

COOH<br />

COOH<br />

COOH<br />

COOH<br />

OH<br />

COOH<br />

HO<br />

COOH<br />

HO 10<br />

10<br />

HO 10<br />

OH<br />

HO<br />

5-series IsoP 12-series IsoP 8-series IsoP 15-series IsoP<br />

OH<br />

HO<br />

COOH<br />

COOH<br />

HO<br />

OH


36 Oxidative Stress and Inflammatory Mechanisms<br />

quantification of F 2-IsoPs has been used to implicate a role for oxidative stress<br />

in the pathophysiology of a number of human conditions and diseases. Notably,<br />

F 2-IsoP levels were shown to be increased in neurodegenerative conditions such<br />

as Alzheimer’s disease, Huntington’s disease, aging, certain types of cancers, and,<br />

of notable importance to consequences of overweight and obesity, atherosclerotic<br />

cardiovascular disease. 14,22–26<br />

A role for oxidant stress in the development and progression of atherosclerosis<br />

has been hypothesized for more than two decades. 27–29 In recent years, however,<br />

the quantification of F 2-IsoPs has allowed investigators to explore, for the first<br />

time, the extent to which humans undergo enhanced oxidant stress under pathophysiological<br />

situations associated with the development of atherosclerotic cardiovascular<br />

disease. These studies have found that increased levels of plasma<br />

and/or urinary F 2-IsoPs are associated with most of the risk factors for atherosclerosis,<br />

including hypercholesterolemia, 30 diabetes mellitus, 31–33 hyperhomocysteinemia,<br />

34 and chronic cigarette smoking. 35–37 This suggests that certain populations<br />

at risk for the disease are under increased oxidant stress.<br />

Oxidant stress levels in the overweight and obese, a population at significant<br />

risk for both atherosclerosis and also for its associated risk factors in which<br />

oxidant stress is increased, had not been studied until recently. Three interesting<br />

studies, however, have independently reported that the overweight and obese<br />

population is under increased oxidant stress.<br />

In the first of these studies, Block and colleagues at the University of California<br />

at Berkeley and Kaiser Permanente in Oakland sought for the first time to<br />

gather large-scale epidemiological data describing oxidative damage that occurs<br />

in normal, healthy populations and the demographic, physical, and nutritional<br />

factors associated with it. 9 More than 300 volunteers (55% women, 45% men)<br />

between the ages of 19 and 80 were recruited, and their complete dietary information<br />

and medical histories were known. Plasma F 2-IsoPs were measured in all<br />

volunteers and statistical analyses were performed to determine whether levels<br />

correlated with race, sex, body mass index (BMI), vitamin intake and levels,<br />

and/or lipid levels. Interestingly, the strongest correlate with increasing levels of<br />

F 2-IsoPs was increasing BMI (Figure 3.2).<br />

In a second study reported in that same year, Davi and colleagues at the<br />

University of Rome reported a smaller study (93 volunteers) in which they<br />

specifically investigated lipid peroxidation in obese women by measuring levels<br />

of urinary F 2-IsoPs. 10 As in the first study, F 2-IsoP levels increased significantly<br />

with increasing BMI. Furthermore, these investigators went on to characterize<br />

the cause-and-effect relationship of this association by examining the effect of<br />

short-term, diet-induced weight loss on urinary F 2-IsoPs. Interestingly, with<br />

weight loss occurred a parallel decrease in BMI and a significant reduction in<br />

F 2-IsoP levels (Figure 3.3). Despite the small size of this weight loss study (12<br />

volunteers), these findings have interesting clinical implications in the possible<br />

role of antioxidants in the treatment of overweight and obese individuals.<br />

In the third and largest study by Keaney et al., urinary F 2-IsoPs were quantified<br />

in nearly 3000 participants in the Framingham Heart Study, and the authors again


Oxidative Stress in Diseases Associated with Overweight and Obesity 37<br />

IsoP (pg/ml)<br />

60<br />

50<br />

40<br />

30<br />


38 Oxidative Stress and Inflammatory Mechanisms<br />

is up-regulated in obesity. 39 Angiotensin II has been shown to induce NADPH<br />

oxidase in various tissues with a resulting increase in superoxide production. 40<br />

Angiotensin II has also been shown to increase LDL uptake by macrophages,<br />

resulting in enhanced lipoprotein oxidation. 40 Further, obesity has been associated<br />

with reduced antioxidant defense mechanisms, including decreased erythrocyte<br />

glutathione and glutathione peroxidase. 41 The extent to which these and other<br />

mechanisms contribute to obesity-associated oxidant stress needs to be explored<br />

and will likely provide key information about the importance of obesity in the<br />

development and progression of disease.<br />

THERAPEUTIC TARGETS FOR REDUCING OXIDANT <strong>STRESS</strong> IN<br />

OVERWEIGHT <strong>AND</strong> OBESE PATIENTS<br />

The findings of Block, 9 Davi, 10 and Keaney 11 are not only important with respect<br />

to the study of basic mechanisms underlying oxidant stress associated with<br />

obesity, but they also have important public health implications in regard to the<br />

treatment of obesity-associated disease. The incidence of overweight and obesity<br />

is becoming more prevalent in the United States and weight loss programs are<br />

often ineffective. 2 Thus, the number of persons with diseases associated with<br />

obesity is going to be a continuing burden to the medical community 42 and novel<br />

strategies to prevent and treat these disorders based on the physiological perturbations<br />

associated with obesity need to be developed and tested. The findings of<br />

the studies discussed herein implicate decreasing in vivo levels of oxidant stress<br />

as one potential therapeutic target for obesity-associated disesase.<br />

ANTIOXIDANTS<br />

One potential therapy to reduce oxidant stress in vivo is antioxidant supplementation.<br />

Animal and human epidemiologic studies carried out in the 1980s and<br />

1990s suggested that antioxidants decreased atherosclerosis, presumably by<br />

reducing oxidant stress. However, prospective clinical trials of antioxidant supplementation<br />

using vitamin E and other agents have been disappointing and<br />

yielded conflicting results. 43–47 Three large trials, ATBC, GISSI, and HOPE,<br />

involving tens of thousands of subjects, failed to show reductions of cardiovascular<br />

events when vitamin E was used at doses ranging from 50 to 400 IU/day. 47–50<br />

On the other hand, two trials, CHAOS and SPACE, involving fewer subjects,<br />

reported near 50% reductions in the incidence of cardiovascular events with<br />

supplementation with 800 IU/day. 51,52<br />

All these studies suffer from the limitation of using only cardiovascular events<br />

as trial endpoints and not assessing oxidant stress in study participants. Thus, it<br />

is impossible to determine whether vitamin E or other antioxidants inhibited<br />

oxidant injuries in the populations studied. Small studies performed by our<br />

laboratory, however, have shown that in order to reduce levels of F 2-IsoPs in vivo,<br />

an individual must take at least 800 IU/day of vitamin E for 16 weeks and probably<br />

more, implying that vitamin E is a very low potency antioxidant. 53 Further,


Oxidative Stress in Diseases Associated with Overweight and Obesity 39<br />

supplementation with that very large amount of vitamin E is not practical due to<br />

safety issues related to consuming that amount of vitamin E. These data coupled<br />

with the clinical trial data suggest that supplementation with vitamin E is unlikely<br />

to prevent atherosclerotic events in humans. However, it is not possible to conclude<br />

from these studies that oxidant stress is not involved in the development<br />

and/or progression of atherosclerosis. Further study is needed to determine<br />

whether decreasing oxidative stress in vivo through antioxidant supplementation<br />

can prevent associated diseases.<br />

ω-3 POLYUNSATURATED FATTY ACIDS<br />

Emerging evidence has implicated increased dietary intake of fish oil containing<br />

large amounts of polyunsaturated fatty eicosapentaenoic acid (20:5, ω-3, EPA)<br />

and docosahexaenoic acid (22:6, ω-3, DHA), as beneficial in the prevention and<br />

treatment of a number of diseases in which environmental and lifestyle factors<br />

play roles. These include atherosclerotic cardiovascular disease and sudden death,<br />

metabolic syndrome, neurodegeneration, and various inflammatory disorders<br />

among others, but the mechanisms by which fish oil is protective are unknown. 54–57<br />

Recent data, however, suggest that the anti-inflammatory effects and other biologically<br />

relevant properties of ω-3 fatty acids are due, in part, to the generation<br />

of various bioactive oxidation products. 58–63<br />

One potentially important anti-atherogenic and anti-inflammatory mechanism<br />

of ω-3 polyunsaturated fatty acids is their interference with the arachidonic acid<br />

cascade that generates pro-inflammatory eicosanoids. 54,64,65 EPA can replace<br />

arachidonic acid in phospholipid bilayers and is also a competitive inhibitor of<br />

COX, reducing the production of 2-series PGs and thromboxane in addition to<br />

4-series leukotrienes. The 3-series-PGs and the 5-series leukotrienes derived from<br />

EPA are either less biologically active or inactive compared to the former products<br />

and are thus considered to exert effects that are less inflammatory. 66,67 Serhan and<br />

colleagues described a group of polyoxygenated DHA and EPA derivatives<br />

termed resolvins that are produced in various tissues. These compounds inhibit<br />

cytokine expression and other inflammatory responses in microglia, skin cells,<br />

and other cell types. 58–61<br />

In addition to studying the enzymatic oxidation of EPA, significant interest<br />

has focused on the biological activities of non-enzymatic free radical-initiated<br />

peroxidation products. Sethi and colleagues reported that EPA oxidized in the<br />

presence of Cu ++ , but not native EPA, significantly inhibits human neutrophil and<br />

monocyte adhesion to endothelial cells — a process linked to the development<br />

of atherosclerosis and other inflammatory disorders. 62,63 This effect was induced<br />

via inhibition of endothelial adhesion receptor expression and was modulated by<br />

the activation of the peroxisome proliferator-activated receptor-α (PPAR-α) by<br />

EPA oxidation products.<br />

Oxidized EPA markedly reduced leukocyte rolling and adhesion to venular<br />

endothelium of lipopolysaccharide-treated mice in vivo and the effect was not<br />

observed in PPAR-α-deficient mice. These studies suggest that the beneficial


40 Oxidative Stress and Inflammatory Mechanisms<br />

effects of fish oil may be mediated, in part, by the anti-inflammatory effects of<br />

oxidized EPA. Similarly, Vallve and colleagues have shown that various nonenzymatically<br />

generated aldehyde oxidation products of EPA (and DHA)<br />

decreased the expression of the CD36 receptor in human macrophages. 68 Upregulation<br />

of this receptor has been linked to atherosclerosis.<br />

Additional recent reports have suggested that other related biological effects<br />

of fish oil, such as modulation of endothelial inflammatory molecules, are related<br />

to the peroxidation products of EPA and DHA. 62,69 Arita and colleagues have also<br />

shown that non-enzymatically oxidized EPA enhances apoptosis in HL-60 leukemia<br />

cells, supporting the contention that oxidized ω-3 polyunsaturated fatty<br />

acids are both anti-proliferative and anti-inflammatory. 70 Similar findings have<br />

been reported in HepG2 (human hepatoma) cells and AH109A (rat liver cancer)<br />

cells. 71,72 Virtually none of these reports, however, have structurally identified the<br />

specific peroxidation products responsible for these effects.<br />

The studies above that attributed some of the beneficial biological activities<br />

of fish oil to the oxidation of its ω-3 fatty acids provided our laboratory with a<br />

rationale to begin studies to systematically define the oxidation of EPA. Specifically,<br />

we were interested in the generation of F-ring IsoP-like compounds (F 3-<br />

IsoPs), structural analogs of the F 2-IsoPs containing an additional double bond<br />

between carbons 17 and 18, based on the hypothesis that these compounds may<br />

contribute to the beneficial biological effects of EPA and fish oil supplementation<br />

in that they may exert anti-inflammatory biological activities compared to F 2-<br />

IsoPs.<br />

One report states that the EPA-derived IsoP, 15-F 3t-IsoP, possesses activity<br />

that is different from 15-F 2t-IsoP in that it does not affect human platelet shape<br />

change or aggregation. 73 Note that 15-F 2t-IsoP is a ligand for the Tx receptor and<br />

induces platelet shape change and also causes vasoconstriction. 15,74 The lack of<br />

activity of 15-F 3t-IsoP is consistent with observations regarding EPA-derived PGs<br />

in that these latter compounds exert either weaker agonist or no effects in comparison<br />

to arachidonate-derived PGs. 55,75.76<br />

Anggard and colleagues provided limited evidence that F 3-IsoPs could be<br />

formed from the oxidation of EPA in vitro. 77 We therefore considered the possibility<br />

that IsoP-like compounds could be formed by the free radical-induced<br />

peroxidation of EPA in vivo. Our studies showed that F 3-IsoPs were generated<br />

from the oxidation of EPA both in vitro and in vivo. 78 Unlike results from the<br />

previous studies by Anggard, the structural characteristics of these compounds<br />

were confirmed by a number of chemical derivatization and mass spectrometric<br />

techniques including LC/ESI/MS/MS.<br />

As expected, six series of F 3-IsoP regioisomers were identified from both in<br />

vitro and in vivo sources. The mass spectrometric fragmentation patterns of these<br />

regioisomers are similar to F 2-IsoP regioisomers and, indeed, information that<br />

we previously acquired with F 2-IsoPs was extremely useful in the characterization<br />

of these molecules. 79 Of note, the relative abundance of 5- and 18-series F 3-IsoPs<br />

predominated over the other series. Such regioisomeric predominance was also<br />

reported for F 2-IsoP regioisomers in which 5- and 15-series compounds were


Oxidative Stress in Diseases Associated with Overweight and Obesity 41<br />

F 2 -IsoPs (ng/g tissue)<br />

75<br />

50<br />

25<br />

0<br />

− EPA + EPA<br />

FIGURE 3.4 Supplementation of mice with eicosapentaenoic acid (EPA) in the diet<br />

reduces endogenous levels of F 2-isoprostanes in the heart.<br />

formed in greater abundance than 8- and 12-series molecules. 79 At least part of<br />

the reason is likely due to the ability of precursors of 8- and 12-series F 2-IsoPs<br />

to undergo further oxidation and cyclization to yield a novel class of compounds<br />

termed dioxolane-endoperoxides. 80 Although undetermined at present, it is likely<br />

that a similar mechanism may account for the predominance of 5- and 18-series<br />

F 3-IsoPs.<br />

Interestingly, the levels of these compounds generated from the oxidation of<br />

EPA significantly exceeded those of F 2-IsoPs generated from arachidonic acid,<br />

perhaps because EPA contains more double bonds and is therefore more easily<br />

oxidizable. Furthermore, in vivo in mice, levels of F 3-IsoPs in tissues such as<br />

heart were virtually undetectable at baseline but supplementation of animals with<br />

EPA markedly increased quantities up to 27.4 ± 5.6 ng/g heart.<br />

Of particular note, we found that EPA supplementation markedly reduced<br />

levels of arachidonate-derived F 2-IsoPs by up to 64% (p


42 Oxidative Stress and Inflammatory Mechanisms<br />

overweight and obesity contribute to disease progression. At the same time, novel<br />

strategies to prevent and treat these disorders based upon our understanding of<br />

the physiological perturbations associated with obesity need to be developed.<br />

The clinical studies reviewed herein by Block, 9 Davi, 10 and Keaney 11 provide<br />

insights into one mechanism, increased oxidant stress, that likely contributes to<br />

the pathophysiology of obesity-associated disease progression. Decreasing<br />

endogenous oxidant stress may therefore be a target for interventions for these<br />

diseases. Interestingly, preliminary animal studies have shown that supplementation<br />

with EPA, one of the major polyunsaturated fatty acids in fish oil, reduces<br />

in vivo levels of F 2-IsoPs, potent mediators of and important biomarkers for<br />

endogenous oxidant stress. Based upon these data and well-known findings that<br />

consumption of fish and fish oil supplementation reduce the incidence of atherosclerosis<br />

and other diseases in humans, fish oil consumption represents a novel<br />

potential treatment to decrease obesity-associated disease.<br />

ACKNOWLEDGMENTS<br />

The work reported herein was supported by National Institutes of Health grants<br />

GM15431, CA77839, DK48831, and ES13125.<br />

REFERENCES<br />

1. http://www.cdc.gov/nchs/products/pubs/pubd/hestats/obese/obse99.htm<br />

2. Eckel, R.H., Barouch, W.W., and Ershow, A.G., Report of the National Heart,<br />

Lung, and Blood Institute–National Institute of Diabetes and Digestive and Kidney<br />

Diseases Working Group on the pathophysiology of obesity-associated cardiovascular<br />

disease, Circulation 105, 2923, 2002.<br />

3. Hedley, A.A. et al., Prevalence of overweight and obesity among U.S. children,<br />

adolescents, and adults, 1999–2002, JAMA 291, 2847, 2004.<br />

4. Bray, G.A., Overweight, mortality and morbidity, in Physical Activity and Obesity,<br />

Bouchard, C. et al., Eds., Human Kinetics Publishers, Champaign, IL, 2000, p. 31.<br />

5. National Task Force on the Prevention and Treatment of Obesity, Overweight,<br />

obesity, and health risk, Arch Int Med 160, 898, 2000.<br />

6. Fontaine, K.R. et al., Years of life lost to obesity, JAMA 289, 187, 2003.<br />

7. Grundy, S.M., Obesity, metabolic syndrome and coronary atherosclerosis, Circulation<br />

105, 2696, 2002.<br />

8. Daniels, S.R., The consequences of childhood overweight and obesity, Future<br />

Child 16, 47, 2006.<br />

9. Block, G. et al., Factors associated with oxidative stress in human populations,<br />

Am J Epidemiol 156, 274, 2002.<br />

10. Davi, G. et al., Platelet activation in obese women: role of inflammation and<br />

oxidant stress, JAMA 288, 2008, 2002.<br />

11. Keaney, J.F., Jr. et al., Obesity and systemic oxidative stress: clinical correlates<br />

of oxidative stress in the Framingham Study, Arterioscler Thromb Vasc Biol 23,<br />

434, 2003.


Oxidative Stress in Diseases Associated with Overweight and Obesity 43<br />

12. Morrow, J.D. and Roberts, L.J., The isoprostanes: unique bioactive products of<br />

lipid peroxidation, Progr Lipid Res 36, 1, 1997.<br />

13. Morrow, J.D. et al., The isoprostanes: unique prostaglandin-like products of free<br />

radical-catalyzed lipid peroxidation, Drug Metab Rev 31, 117, 1999.<br />

14. Famm, S.S. and Morrow, J.D., The isoprostanes: unique products of arachidonic<br />

acid oxidation: a review, Curr Med Chem 10, 1723, 2003.<br />

15. Morrow, J.D. et al., A series of prostaglandin F2-like compounds are produced in<br />

vivo in humans by a non-cyclooxygenase, free radical-catalyzed mechanism, Proc<br />

Natl Acad Sci USA 87, 9383, 1990.<br />

16. Roberts, L.J. and Morrow, J.D., Measurement of F 2-isoprostanes: a reliable index<br />

of oxidant stress in vivo, Free Radical Biol Med 28, 505, 2000.<br />

17. Morrow, J.D. and Roberts, L.J., Mass spectrometric quantification of F 2-isoprostanes<br />

in biological fluids and tissues as a measure of oxidant stress, Meth Enzymol<br />

300, 3, 1999.<br />

18. Morrow, J.D., The isoprostanes: their quantification as an index of oxidant stress<br />

status in vivo, Drug Metab Rev 32, 377, 2000.<br />

19. Griffiths, H.R. et al., Biomarkers, Mol Aspects Med 23, 101, 2002.<br />

20. Kadiiska, M.B. et al., Biomarkers of oxidative stress study II: are oxidation<br />

products of lipids, proteins, and DNA markers of CCl 4 poisoning? Free Radical<br />

Biol Med 38, 698, 2005.<br />

21. Kadiiska, M.B. et al., Biomarkers of oxidative stress study III: effects of the<br />

nonsteroidal anti-inflammatory agents indomethacin and meclofenamic acid on<br />

measurements of oxidative products of lipids in CCl 4 poisoning, Free Radical<br />

Biol Med 38, 711, 2005.<br />

22. Gniwotta, C. et al., Prostaglandin F2-like compounds, F2-isoprostanes, are present<br />

in increased amounts in human atherosclerotic lesions, Arterioscler Thromb Vasc<br />

Biol 17, 3236, 1997.<br />

23. Gopaul, N.K. et al., Formation of F2-isoprostanes during aortic endothelial cellmediated<br />

oxidation of low density lipoprotein, FEBS Lett 348, 297, 1994.<br />

24. Montine, K.S. et al., Isoprostanes and related products of lipid peroxidation in<br />

neurodegenerative diseases, Chem Phys Lipids 128, 117, 2004.<br />

25. Montine, T.J. et al., Cerebrospinal fluid F2-isoprostanes are elevated in Huntington's<br />

disease, Neurology 52, 1104, 1999.<br />

26. Montine, T.J. et al., Cerebrospinal fluid F2-isoprostane levels are increased in<br />

Alzheimer's disease, Ann Neurol 44, 410, 1998.<br />

27. Heinecke, J.W., Oxidants and antioxidants in the pathogenesis of atherosclerosis:<br />

implications for the oxidized low density lipoprotein hypothesis, Atherosclerosis<br />

141, 1, 1998.<br />

28. Berliner, J. et al., Oxidized lipids in atherogenesis: formation, destruction and<br />

action, Thromb Haemost 78, 195, 1997.<br />

29. Chisholm, G.M. and Steinberg, D., The oxidative modification hypothesis of<br />

atherogenesis: an overview, Free Radical Biol Med 18, 1815, 2000.<br />

30. Davi, G. et al., In vivo formation of 8-epi-prostaglandin F2 alpha is increased in<br />

hypercholesterolemia, Arterioscler Thromb Vasc Biol 17, 3230, 1997.<br />

31. Davi, G. et al., In vivo formation of 8-iso-prostaglandin F2 alpha and platelet<br />

activation in diabetes mellitus: effects of improved metabolic control and vitamin<br />

E supplementation, Circulation 99, 224, 1999.


44 Oxidative Stress and Inflammatory Mechanisms<br />

32. Davi, G. et al., Enhanced lipid peroxidation and platelet activation in the early<br />

phase of type 1 diabetes mellitus: role of interleukin-6 and disease duration,<br />

Circulation 107, 3199, 2003.<br />

33. Gopaul, N.K. et al., Plasma 8-epi-PGF2 alpha levels are elevated in individuals<br />

with non-insulin dependent diabetes mellitus, FEBS Lett 368, 225, 1995.<br />

34. Voutilainen, S. et al., Enhanced in vivo lipid peroxidation at elevated plasma total<br />

homocysteine levels, Arterioscler Thromb Vasc Biol 19, 1263, 1999.<br />

35. Frei, B. et al., Gas phase oxidants of cigarette smoke induce lipid peroxidation<br />

and changes in lipoprotein properties in human blood plasma: protective effects<br />

of ascorbic acid, Biochem J 277, 133, 1991.<br />

36. Morrow, J.D. et al., Increase in circulating products of lipid peroxidation (F2isoprostanes)<br />

in smokers: smoking as a cause of oxidative damage, New Engl J<br />

Med 332, 1198, 1995.<br />

37. Obwegeser, R. et al., Maternal cigarette smoking increases F2-isoprostanes and<br />

reduces prostacyclin and nitric oxide in umbilical vessels, Prostaglandins Other<br />

Lipid Mediat 57, 269, 1999.<br />

38. Chan, J.C. et al., The central roles of obesity-associated dyslipidaemia, endothelial<br />

activation and cytokines in the metabolic syndrome: analysis by structural equation<br />

modelling, Int J Obes Relat Metab Disord 26, 994, 2002.<br />

39. Barton, M. et al., Obesity is associated with tissue-specific activation of renal<br />

angiotensin-converting enzyme in vivo: evidence for a regulatory role of endothelin,<br />

Hypertension 35, 329, 2000.<br />

40. Brasier, A.R., Recinos, A., 3rd, and Eledrisi, M.S., Vascular inflammation and the<br />

renin-angiotensin system, Arterioscler Thromb Vasc Biol 22, 1257, 2002.<br />

41. Trevisan, M. et al., Correlates of markers of oxidative status in the general population,<br />

Am J Epidemiol 154, 348, 2001.<br />

42. Manson, J.E. and Bassuk, S.S., Obesity in the United States: a fresh look at its<br />

high toll, JAMA 289, 229, 2003.<br />

43. Rimm, E.B. et al., Vitamin E consumption and the risk of coronary heart disease<br />

in men, New Engl J Med 328, 1450, 1993.<br />

44. Stampfer, M.J. et al., Vitamin E consumption and the risk of coronary disease in<br />

women, New Engl J Med 328, 1444, 1993.<br />

45. Diaz, M.N. et al., Antioxidants and atherosclerotic heart disease, New Engl J Med<br />

337, 408, 1997.<br />

46. Pratico, D. et al., Vitamin E suppresses isoprostane generation in vivo and reduces<br />

atherosclerosis in ApoE-deficient mice, Nat Med 4, 1189, 1998.<br />

47. Heinecke, J.W., Is the emperor wearing clothes? Clinical trials of vitamin E and<br />

the LDL oxidation hypothesis, Arterioscler Thromb Vasc Biol 21, 1261, 2001.<br />

48. Alpha-Tocopherol–Beta Carotene Cancer Prevention Study Group, The effect of<br />

vitamin E and beta carotene on the incidence of lung cancer and other cancers in<br />

male smokers, New Engl J Med 330, 1029, 1994.<br />

49. Gruppo Italiano per lo Studio della Sopravvivenza nell'Infarto Miocardico, Dietary<br />

supplementation with n-3 polyunsaturated fatty acids and vitamin E after myocardial<br />

infarction: results of the GISSI Prevenzione Trial, Lancet 354, 447, 1999.<br />

50. Yusuf, S. et al., Vitamin E supplementation and cardiovascular events in high-risk<br />

patients, New Engl J Med 342, 154, 2000.<br />

51. Stephens, N.G. et al., Randomised controlled trial of vitamin E in patients with<br />

coronary disease: Cambridge Heart Antioxidant Study (CHAOS), Lancet 347,<br />

781, 1996.


Oxidative Stress in Diseases Associated with Overweight and Obesity 45<br />

52. Boaz, M. et al., Secondary prevention with antioxidants of cardiovascular disease<br />

in endstage renal disease (SPACE): randomised placebo-controlled trial, Lancet<br />

356, 1213, 2000.<br />

53. Morrow, J.D. et al., Alpha-tocopherol supplementation reduces plasma isoprostane<br />

levels in hypercholesterolemic humans only at doses of 800 IU or greater, Circulation<br />

106, 165, 2002.<br />

54. Kris-Etherton, P.M., Harris, W.S., and Appel, L.J., Fish consumption, fish oil,<br />

omega-3 fatty acids, and cardiovascular disease, Arterioscler Thromb Vasc Biol<br />

23, 20, 2003.<br />

55. Kris-Etherton, P.M., Harris, W.S., and Appel, L.J., Omega-3 fatty acids and cardiovascular<br />

disease: new recommendations from the American Heart Association,<br />

Arterioscler Thromb Vasc Biol 23, 151, 2003.<br />

56. Ruxton, C., Health benefits of omega-3 fatty acids, Nurs Stand, 18, 38, 2004.<br />

57. Ruxton, C.H. et al., The health benefits of omega-3 polyunsaturated fatty acids:<br />

a review of the evidence, J Hum Nutr Diet 17, 449, 2004.<br />

58. Hong, S. et al., Novel docosatrienes and 17S-resolvins generated from docosahexaenoic<br />

acid in murine brain, human blood, and glial cells: autacoids in antiinflammation,<br />

J Biol Chem 278, 14677, 2003.<br />

59. Serhan, C.N. and Levy, B., Novel pathways and endogenous mediators in antiinflammation<br />

and resolution, Chem Immunol Allergy 83, 115, 2003.<br />

60. Serhan, C.N. et al., Novel functional sets of lipid-derived mediators with antiinflammatory<br />

actions generated from omega-3 fatty acids via cyclooxygenase 2nonsteroidal<br />

antiinflammatory drugs and transcellular processing, J Exp Med 192,<br />

1197, 2000.<br />

61. Serhan, C.N. et al., Resolvins: a family of bioactive products of omega-3 fatty<br />

acid transformation circuits initiated by aspirin treatment that counter proinflammation<br />

signals, J Exp Med 196, 1025, 2002.<br />

62. Sethi, S., Eastman, A.Y., and Eaton, J.W., Inhibition of phagocyte–endothelium<br />

interactions by oxidized fatty acids: a natural anti-inflammatory mechanism? J<br />

Lab Clin Med 128, 27, 1996.<br />

63. Sethi, S., Inhibition of leukocyte-endothelial interactions by oxidized omega-3<br />

fatty acids: a novel mechanism for the anti-inflammatory effects of omega-3 fatty<br />

acids in fish oil, Redox Rep 7, 369, 2002.<br />

64. Uauy, R., Mena, P., and Valenzuela, A., Essential fatty acids as determinants of<br />

lipid requirements in infants, children and adults, Eur J Clin Nutr 53, Suppl 1,<br />

S66, 1999.<br />

65. Smith, W.L., Cyclooxygenases, peroxide tone, and the allure of fish oil, Curr Opin<br />

Cell Biol 17, 174, 2005.<br />

66. Calder, P.C., Polyunsaturated fatty acids, inflammation, and immunity, Lipids 36,<br />

1007, 2001.<br />

67. Yang, P. et al., Formation and antiproliferative effect of prostaglandin E(3) from<br />

eicosapentaenoic acid in human lung cancer cells, J Lipid Res 45, 1030, 2004.<br />

68. Vallve, J.C. et al., Unsaturated fatty acids and their oxidation products stimulate<br />

CD36 gene expression in human macrophages, Atherosclerosis 164, 45, 2002.<br />

69. Das, U.N., Essential fatty acids, lipid peroxidation, and apoptosis, Prostaglandins<br />

Leukotr Essent Fatty Acids 61, 157, 1999.<br />

70. Arita, K. et al., Mechanisms of enhanced apoptosis in HL-60 cells by UV-irradiated<br />

n-3 and n-6 polyunsaturated fatty acids, Free Radic Biol Med 35, 189, 2003.


46 Oxidative Stress and Inflammatory Mechanisms<br />

71. Kokura, S. et al., Enhancement of lipid peroxidation and of the antitumor effect<br />

of hyperthermia upon combination with oral eicosapentaenoic acid, Cancer Lett<br />

185, 139, 2002.<br />

72. Kiserud, C.E., Kierulf, P., and Hostmark, A.T., Effects of various fatty acids alone<br />

or combined with vitamin E on cell growth and fibrinogen concentration in the<br />

medium of HepG2 cells, Thromb Res 80, 75, 1995.<br />

73. Pratico, D. et al., Local amplification of platelet function by 8-epi prostaglandin<br />

F2-alpha is not mediated by thromboxane receptor isoforms, J Biol Chem 271,<br />

14916, 1996.<br />

74. Morrow, J.D., Minton, T.A., and Roberts, L.J., 2nd, The F2-isoprostane, 8-epiprostaglandin<br />

F2 alpha, a potent agonist of the vascular thromboxane/endoperoxide<br />

receptor, is a platelet thromboxane/endoperoxide receptor antagonist, Prostaglandins<br />

44, 155, 1992.<br />

75. Kulkarni, P.S. and Srinivasan, B.D., Prostaglandins E3 and D3 lower intraocular<br />

pressure, Invest Ophthalmol Vis Sci 26, 1178, 1985.<br />

76. Balapure, A.K. et al., Structural requirements for prostaglandin analog interaction<br />

with the ovine corpus luteum prostaglandin F2 alpha receptor: implications for<br />

development of a photoaffinity probe, Biochem Pharmacol 38, 2375, 1989.<br />

77. Nourooz-Zadeh, J., Halliwell, B., and Anggard, E.E., Evidence for the formation<br />

of F3-isoprostanes during peroxidation of eicosapentaenoic acid, Biochem Biophys<br />

Res Commun 236, 467, 1997.<br />

78. Gao, L. et al., Formation of F-ring isoprostane-like compounds (F 3-isoprostanes)<br />

in vivo from eicosapentaenoic acid, J Biol Chem 281, 14092, 2006.<br />

79. Waugh, R.J. et al., Identification and relative quantitation of F2-isoprostane regioisomers<br />

formed in vivo in the rat, Free Radic Biol Med 23, 943, 1997.<br />

80. Yin, H., Morrow, J.D., and Porter, N.A., Identification of a novel class of endoperoxides<br />

from arachidonate autoxidation, J Biol Chem 279, 3766, 2004.


4<br />

CONTENTS<br />

Metabolic Syndrome<br />

Due to Early Life<br />

Nutritional<br />

Modifications<br />

Malathi Srinivasan, Paul Mitrani, and<br />

Mulchand S. Patel<br />

Introduction .........................................................................................................47<br />

Role of Metabolic Programming in Etiology of Obesity Epidemic..................49<br />

Pre-Natal Metabolic Programming ...........................................................50<br />

Metabolic Programming Due to Altered Dietary Experience in<br />

Immediate Post-Natal Period ........................................................53<br />

High Carbohydrate (HC) Rat Model ........................................................54<br />

Metabolic Programming in HC Rats: Immediate Effects ........................56<br />

Persistent Effects in Adulthood.................................................................58<br />

Generational Effects ..................................................................................59<br />

Correlation of Altered Nutritional Experience in Early Life to<br />

Subsequent High Incidence of Obesity and Metabolic Syndrome ..........62<br />

Acknowledgment.................................................................................................63<br />

References ...........................................................................................................63<br />

INTRODUCTION<br />

Obesity is a rapidly growing worldwide epidemic that is especially prevalent in<br />

the United States. Recent data estimate that 66% of American adults are overweight,<br />

with about 32% being obese. 1 Interestingly, dramatic increases in obesity<br />

rates are relatively recent phenomena, with a rapid rise beginning in the 1980s. 2<br />

Between 1960 and 1980, the prevalence of obesity in the United States increased<br />

less than 2% (from 13 to 15%) while obesity rates in the last 25 years have more<br />

than doubled. 3,4 Similar trends are seen internationally, indicating that the obesity<br />

epidemic is a global health problem. 5<br />

47


48 Oxidative Stress and Inflammatory Mechanisms<br />

Because obesity is associated with increased risks for the development of<br />

diabetes, cardiovascular disease, dyslipidemia, and hypertension, the epidemic of<br />

obesity will be followed by an epidemic of these diseases. 5,6 Therefore, it is no<br />

surprise that in parallel with the obesity epidemic an explosion in the incidence<br />

of type 2 diabetes mellitus (T2DM) has ensued. T2DM is one of the biggest<br />

health concerns related to obesity, affecting more than 18 million people in the<br />

United States and ranks as the sixth leading cause of death. 7 Between 1990 and<br />

1998, the prevalence of diabetes increased by 33% in the United States. 8<br />

Metabolic syndrome, previously referred to as syndrome X, 9 is defined by<br />

the clustering of cardiovascular risk factors including obesity, diabetes, hypertension,<br />

and dyslipidemia, 5,9 and is associated with insulin resistance. 5 Metabolic<br />

syndrome is now estimated to affect ~34% of American adults and up to<br />

36% of adult Europeans. 10 The American Medical Association estimates that<br />

the medical costs associated with the treatment of obesity-associated diseases<br />

such as T2DM and cardiovascular disorders now exceed the costs of smokingrelated<br />

diseases. 11<br />

The adult obesity epidemic has been accompanied by an analogous epidemic<br />

of childhood obesity, 12 especially in the United States. 13 While the prevalence of<br />

obesity in children has been increasing since the 1960s, the last 25 years have<br />

shown a rapid acceleration in childhood obesity rates. 13 During this period, the<br />

proportion of overweight children has almost doubled, while the proportion of<br />

overweight adolescents has nearly tripled, 7 leading to the increased incidence of<br />

T2DM in this sector of the population in the United States. 14 Additionally, the<br />

increase in the incidence of overweight children places them at a greater risk of<br />

developing obesity and T2DM as adults. 15 If the epidemics of obesity, diabetes,<br />

and metabolic syndrome continue to rise, the prognosis of worldwide health will<br />

continue to decline. Hence there is a sense of urgency in the need to further our<br />

knowledge about the etiology of obesity, specifically related to the steep increase<br />

in its incidence within the past 25 years, in an effort to curb the surge in the<br />

occurrence of obesity-enhanced diseases.<br />

The rapid rise in obesity rates since the 1980s suggests that the cause is a<br />

combination of various behavioral and environmental influences rather than<br />

specific genetic and biological changes. 1 It is commonly believed that the recent<br />

trends in obesity are the results of increased consumption of calorie-dense, high<br />

fat, and high carbohydrate foods along with decreased physical activity. 7<br />

Although these factors do contribute to the onset of obesity, it is now recognized<br />

that obesity also results from programmed impairments of energy balance that<br />

increase vulnerability to negative environments that include poor diet and<br />

limited exercise. 1<br />

Recent evidence suggests that both increases in food consumption and<br />

changes in the quality of nutrition may play decisive roles. For most of the 20th<br />

century, caloric consumption remained virtually constant, but then began to rise<br />

in the 1980s due to increased consumption of fats and sugars. 2,16 Of particular<br />

interest is the finding that the consumption of carbohydrates as a percentage of<br />

total calorie intake (due to increases in the consumption of soft drinks and fruit


Metabolic Syndrome Due to Early Life Nutritional Modifications 49<br />

juices) increased by approximately 20% over the past few decades. 2 In addition,<br />

evidence indicates that the quality of consumed carbohydrates is changing as<br />

high glycemic simple sugars are replacing low glycemic complex carbohydrates. 13<br />

Compared to high glycemic diets, low glycemic diets have been found to increase<br />

satiety and delay return of hunger, lower post-prandial insulin levels, and increase<br />

insulin sensitivity. 13<br />

These studies indicate that changes in the quantity and quality of dietary<br />

carbohydrate are important components in the pathogenesis of obesity and metabolic<br />

syndrome. However, in addition to changing dietary habits in adulthood,<br />

recent evidence indicates that exposure to altered nutrition during very early<br />

phases of life (pre-natal and immediate post-natal) can result in metabolic programming<br />

that predisposes individuals to the development of metabolic syndrome<br />

(Figure 4.1).<br />

ROLE OF METABOLIC PROGRAMMING IN ETIOLOGY OF<br />

OBESITY EPIDEMIC<br />

Population-based evidence and studies of early nutritional experiences in animals<br />

suggest that different nutritional insults during fetal or neonatal life may result<br />

in increased risks of developing metabolic diseases such as obesity and metabolic<br />

syndrome later in life. 10 Metabolic programming is a phenomenon in which a<br />

stimulus or insult that occurs during a critical period of organogenesis in early<br />

life results in permanent alterations in the structures and functions of affected<br />

organs and increased susceptibility to adult disease (Figure 4.1). 17,18<br />

The fetal origins hypothesis originally proposed by Barker posits that metabolic<br />

programming occurs in any situation in which a stimulus or insult during<br />

development establishes a permanent physiological response. 19 A related theory<br />

called the thrifty phenotype hypothesis proposes that a poor nutritional environment<br />

in utero results in fetal metabolic programming that maximizes uptake and<br />

conservation of available nutrients. 20,21<br />

It is well established that critical periods of development that coincide with<br />

periods of rapid cell division are important for the abilities of specific tissues and<br />

organs to differentiate and mature in preparation for survival after birth. 22,23<br />

Therefore, poor fetal nutrition results in selective protection of the brain at the<br />

expense of other organs such as the pancreas and liver, and results in altered<br />

growth and permanent changes in organ structure and function (Figure 4.1). 22,24<br />

Metabolic programming is meant to confer an adaptive advantage to an infant<br />

exposed to a post-natal deficiency similar to the one encountered in utero. This<br />

is also related to the idea of thrifty genes that are involved in the adaptive response<br />

to environments with scarce availability of food and function to increase survival<br />

in such environments. However, when these genes are introduced to situations<br />

where resources are abundant, they begin to lose their adaptive advantages and<br />

become detrimental due to increased sensitivity to and utilization of those<br />

resources. 20,25


50 Oxidative Stress and Inflammatory Mechanisms<br />

FIGURE 4.1 Metabolic programming due to early life nutritional challenges. Pre-natal<br />

programming can be caused by maternal malnutrition, gestational diabetes, or by an<br />

overweight/insulin resistant pregnancy. Post-natal programming can occur due to altered<br />

neonatal nutrition (under- or over-nutrition, caloric redistribution). The maladaptations in<br />

target organs during the period of the nutritional modification persist in later life, resulting<br />

in increased risk for metabolic diseases in adulthood.<br />

PRE-NATAL METABOLIC PROGRAMMING<br />

Prenatal<br />

Programming<br />

Poor fetal nutrition<br />

- maternal malnutrition, protein deficiency, high-fat<br />

Gestational Diabetes<br />

Impairments of fetal growth,<br />

pancreatic development,<br />

and hypothalamic development<br />

Hormonal/metabolic abnormalities<br />

Metabolic Diseases in Adulthood<br />

(obesity, T2DM, metabolic syndrome)<br />

Impairments of neonatal growth,<br />

pancreatic development,<br />

and hypothalamic development<br />

Hormonal/metabolic abnormalities<br />

Inadequate neonatal nutrition<br />

- maternal malnutrition during lactation<br />

- HC dietary intervention = altered calorie sources<br />

Excessive neonatal nutrition<br />

- overfeeding (due to decreased litter sizes)<br />

Postnatal<br />

Programming<br />

Original studies by Barker et al. showed that adverse intrauterine conditions such<br />

as under-nutrition during pregnancy caused disproportionate fetal growth, resulting<br />

in low birth weight and adult T2DM 18,26 as well as hypertension and dyslipidemia.<br />

26,28 Data from several other epidemiological studies carried out in different


Metabolic Syndrome Due to Early Life Nutritional Modifications 51<br />

parts of the world support this association between fetal development in an<br />

undernourished maternal environment and the later onset of metabolic diseases. 29<br />

Although data from human epidemiological studies have indicated the connection<br />

between impaired fetal development and adult onset diseases such as<br />

T2DM, hypertension, cardiovascular problems, and other conditions, intrinsic<br />

difficulties surround the conduct of long-term studies in human populations.<br />

Therefore, several animal models have been developed to mimic the human<br />

situation to further our knowledge on the role of maternal nutritional status during<br />

pregnancy and the long-term outcome for progeny. Significant among these<br />

models are maternal protein or total caloric restriction and gestational diabetes<br />

(Figure 4.1). Several elegant reviews summarize these models in detail. 30–33<br />

TABLE 4.1<br />

Summary of Immediate and Long-Term Consequences for Progeny Due<br />

to Altered Maternal (Pregnancy and Lactation) Nutritional Experience<br />

I. Dietary protein restriction during gestation and lactation<br />

Immediate effects<br />

Fetal growth retardation<br />

Impairment in pancreatic islet structure and function<br />

Malformation of hypothalamic nuclei<br />

Persistent effects<br />

Increased insulin sensitivity in young adults<br />

Altered insulin signaling in peripheral tissues<br />

Renal defects and hypertension<br />

Development of glucose intolerance with advancing age<br />

II. Total caloric restriction during gestation and lactation<br />

30% caloric restriction<br />

Increased fasting insulin, hyperphagia, adult onset obesity and hypertension<br />

50% caloric restriction<br />

Impaired islet development, glucose intolerance<br />

III. Diabetes during pregnancy<br />

Mild diabetes<br />

Impaired development of fetal islets<br />

Impaired glucose tolerance and reduced insulin secretory capacity in adulthood<br />

Severe diabetes<br />

Fetal growth retardation resulting in smaller-sized adults<br />

Over-stimulation of fetal β cells due to fetal hyperglycemia<br />

Onset of insulin resistance in liver, muscle, and adipose tissue in adulthood<br />

IV. Altered intrauterine environment (maternal hyperinsulinemia/obesity) due to high<br />

fat diet consumption<br />

Fetal hyperinsulinemia and altered insulin secretory capacity<br />

Alterations in β and α cell volume and number on post-natal day 1<br />

Abnormal glucose homeostasis, impaired insulin secretion, increased body adiposity,<br />

hyperlipidemia, altered endothelial function and pro-atherogenic lesions in adulthood


52 Oxidative Stress and Inflammatory Mechanisms<br />

Maternal protein malnutrition due to a low protein diet results in low protein<br />

(LP) progeny that are underweight, hypophagic, hypoglycemic, and hypoinsulinemic<br />

throughout life (Table 4.1) 34–36 and show significant alterations in pancreatic islets,<br />

hypothalamus, liver, and muscle (Table 4.1). 37–40 Interestingly, offspring of proteinmalnourished<br />

rats showed reductions in the relative weights of the pancreas, muscle,<br />

and liver, while the weights of the brain and lungs were protected 22,41 — consistent<br />

with the idea that nutrients are diverted to essential organs when scarce.<br />

While hypoinsulinemia in LP progeny is attributable to lower blood glucose<br />

levels, evidence indicates that insufficient β cell stimulation by amino acids during<br />

fetal development impairs insulin secretory capacity after birth. 42 Protein malnutrition<br />

during pregnancy reduces maternal plasma amino acid concentrations 43<br />

and thus placental transfer of amino acids to the fetus. 42 Reduced fetal levels of<br />

taurine have been associated with impaired development of normal fetal insulin<br />

secretion, 22,40,44,45 while supplemental taurine given to mothers on LP diets was<br />

able to normalize insulin secretion in fetal islets. 22,46<br />

Reduced insulin secretion by LP progeny is associated with reduced islet<br />

proliferation, size, insulin content, and vascularization, 22,34,35,44 as well as a permanent<br />

reduction in the number of pancreatic β cells. 35 Neonatal rats showed<br />

similar defects in pancreatic development when mothers were kept on low protein<br />

diets during lactation. 22,34 While reduced insulin secretion may confer an adaptive<br />

response for survival in a protein-restricted environment, the resulting impairment<br />

of insulin secretory capacity increases adulthood risk of glucose intolerance and<br />

diabetes. 22 Therefore, abnormal stimulation of β cells in early life causes malprogramming<br />

of β cell function that results in impaired insulin secretion throughout<br />

life and may increase the risk of T2DM.<br />

Maternal protein restriction during gestation and lactation results in offspring<br />

(LP) with altered programming not only of pancreatic islets but also of brain<br />

areas involved in the regulation of energy homeostasis and increases the risk of<br />

adult metabolic diseases, including decreased insulin secretion and increased<br />

sympathetic nervous system activity in association with hypertension. 36,47<br />

Among the affected brain regions are the ventromedial nucleus of the hypothalamus<br />

and paraventricular nucleus, which show greater relative volumes, along<br />

with increases in neuronal density in LP rats compared to control animals. 36 This<br />

increase in neuronal density is associated with increases in the activities of the<br />

ventromedial nucleus and paraventricular nucleus, which is consistent with data<br />

showing that ventromedial nucleus activity inhibits insulin secretion and stimulates<br />

sympathetic nervous system activity, 48 while activity of the paraventricular<br />

nucleus stimulates increases in blood pressure. 36,49 In addition, LP rats showed<br />

reduced numbers of arcuate neuropeptide Y neurons that may contribute to alterations<br />

in neuropeptide Y-mediated regulation of energy homeostasis including<br />

sympathetic nervous system activity. 36<br />

In the case of total caloric restriction, 50% reduction of food availability<br />

during the second and third weeks of pregnancy in rats resulted in impaired<br />

pancreatic development and glucose intolerance in the progeny at 1 year of age<br />

(Table 4.1). 50 Severe caloric restriction (access to only 30% of the food consumed


Metabolic Syndrome Due to Early Life Nutritional Modifications 53<br />

by controls) resulted in increased fasting plasma insulin and leptin, hyperphagia,<br />

hypertension, and obesity in the adult progeny. 51<br />

Diabetic pregnancies are associated with significant structural and functional<br />

changes in the fetal endocrine pancreas. For example, a mild diabetic pregnancy in<br />

rats resulted in hypertrophy and hyperplasia of fetal islets, 52 as well as enhanced<br />

proliferative capacity (Table 4.1). 53,54 In addition, fetal islets have increased insulin<br />

content 55 and increased responsiveness to glucose-stimulated insulin secretion. 55,56 A<br />

severely diabetic pregnancy, on the other hand, resulted in fetal islets that had lower<br />

insulin content 54 and reduced glucose-stimulated insulin secretion (Table 4.1). 56 Further,<br />

offspring of rat mothers with gestational diabetes showed significant increases<br />

in the density of arcuate neuropeptide Y neurons associated with decreased sympathetic<br />

nervous system activity and increased weight gain. 36,57<br />

Although several studies have focused on the consequences of a malnourished<br />

pregnancy that were important during periods of famine and war, the current dietary<br />

habits of people (consumption of high energy foods) especially in Westernized<br />

societies require investigations into the consequences for the progeny arising from<br />

such dietary practices in women. It has been suggested that the consumption of<br />

energy-dense foods is a contributing factor in the etiology of the current obesity<br />

epidemic and may well be responsible for a significant and increasing proportion of<br />

women being overweight during pregnancy. 58 Chronic feeding of a high fat diet to<br />

female rats bears a close resemblance to the current dietary habits of humans in<br />

Western societies and hence has physiological relevance to the human situation.<br />

Long-term consumption of a high fat diet resulted in hyperinsulinemia and increased<br />

insulin secretory responses to various secretogogues in term fetuses (Table 4.1). 59<br />

Additionally, the adult progeny of such mothers were obese, glucose-intolerant,<br />

hyperinsulinemic, and exhibited abnormal insulin secretory responses to glucose. 59<br />

Cerf et al. 60 demonstrated that feeding a high fat diet to female rats throughout<br />

gestation alone resulted in significant decreases in β cell volume and number and<br />

converse changes in α cells, resulting in hyperglycemia in 1-day-old newborn rat<br />

pups without changes in serum insulin concentrations. Several reports indicate<br />

the long-term consequences of a high fat diet during gestation only or during<br />

both gestation and lactation. These include abnormal glucose homeostasis,<br />

reduced whole body insulin sensitivity, impaired β cell insulin secretion, and<br />

changes in the structure of the pancreas, 61,62 defective mesenteric artery endothelial<br />

function, 63 hypertension, 64 alterations in conduit artery and renal functions, 65<br />

increased body adiposity, 61,63 deranged blood lipid profile, 61,64 hyperleptinemia, 63<br />

and pro-artherogenic lesions 66 in adult progeny (Table 4.1). Kozak et al. 67 demonstrated<br />

that a high fat diet during gestation and lactation affected body weight<br />

regulation in adult progeny via alterations in the functioning of neuropeptide Y.<br />

METABOLIC PROGRAMMING DUE TO ALTERED DIETARY EXPERIENCE IN<br />

IMMEDIATE POST-NATAL PERIOD<br />

Critical windows of development of organs and metabolic processes also extend<br />

into the immediate post-natal period, suggesting vulnerability of this period to


54 Oxidative Stress and Inflammatory Mechanisms<br />

metabolic programming effects via altered nutritional experiences. In the case of<br />

mammals, natural rearing by mothers via breast feeding is the nature-made<br />

program for rearing of newborns. In humans it is well established that breast<br />

feeding is the optimal form of infant nutrition because it confers immunologic,<br />

psychological, and developmental benefits to the infant. 68,69 Several studies have<br />

indicated that breast feeding is protective against the development of obesity,<br />

diabetes, cardiovascular disease, and metabolic syndrome, 70,71 and this protection<br />

directly correlated with the duration of breast feeding. 72<br />

The association of breast feeding with decreased prevalence of childhood<br />

obesity 73 is thought to involve protection against metabolic disease through careful<br />

regulation of an infant’s calorie intake, 74 insulin secretion, 75 and adiposity. 72<br />

It also may help program the complex circuitry involved in the regulation of<br />

energy homeostasis that persists throughout life. 72<br />

McCance was the first researcher to demonstrate from studies in rats that<br />

changing the availability of milk during the suckling period had permanent effects<br />

on growth trajectories. 76 These studies provided the first evidence indicating a<br />

connection between nutritional experiences during the suckling period and longterm<br />

metabolic effects. Further studies stemming from adjustment of litter size<br />

in rats indicated that rats raised in small litters demonstrated increased body<br />

weight gain during the suckling period that persisted into the post-weaning period<br />

accompanied by hyperphagia, hyperleptinemia, hyperinsulinemia, and permanent<br />

alterations in gene expression and insulin secretory capacity in islets. 57,77–79<br />

Adult onset obesity in rats raised in small litters is supported by altered<br />

programming of the energy circuitry in the hypothalamus (Table 4.2). 80 Post-natal<br />

over-nutrition has been shown to alter the functions of the arcuate nucleus and<br />

ventromedial hypothalamus. For example, overfed rats showed increases in the<br />

density of arcuate neuropeptide Y neurons in association with increases in body<br />

weight and food intake. 81,82 In addition, the direction of responses to corticotropin<br />

releasing factor is apparently reversed in overfed rats. 83 Also in overfed rats (small<br />

litter size), increased inhibition of the ventromedial hypothalamic neurons by<br />

agouti-related peptide has been reported. 84 Rats raised in large litters during the<br />

suckling period maintained diminished growth patterns throughout life and<br />

showed reduced plasma insulin levels compared to rats from small litters of three<br />

or four pups per dam (Table 4.2). 76,85<br />

HIGH CARBOHYDRATE (HC) RAT MODEL<br />

Most studies on metabolic programming using animal models are involved with<br />

an altered intrauterine environment induced by alterations in the nutritional status<br />

of the mother during pregnancy. Due to difficulties in rearing newborn pups away<br />

from their dams, very few studies have examined the effects of nutritional modifications<br />

during the suckling period. Adjustment of litter size is an approach in<br />

this direction but results only in a decrease or increase in the availability of total<br />

calories during the suckling period and does not permit investigation of alterations<br />

in the quality of nutrition without affecting total caloric intake. In the rat, the


Metabolic Syndrome Due to Early Life Nutritional Modifications 55<br />

TABLE 4.2<br />

Summary of Immediate and Long-Term Consequences of Altered<br />

Nutritional Experience during Suckling Period<br />

I. Under- or over-nourishment due to adjustment of litter size<br />

Large-sized litters<br />

Reduced growth, alterations in islet functions<br />

Small-sized litters<br />

Increased growth, hyperleptinemia, and hyperinsulinemia from the start and persisting<br />

into adulthood<br />

Molecular and functional alterations in adult islets<br />

Alterations in hypothalamic neuropeptide expression and function<br />

II. Caloric redistribution (HC dietary intervention) in suckling period<br />

Immediate effects<br />

Immediate onset of hyperinsulinemia with euglycemia<br />

Alterations in islet insulin secretory capacity such as increased response to lower glucose<br />

concentrations, modified responses to incretins and neuroendocrine effectors, and ability<br />

to secrete moderate amounts of insulin in absence of glucose and calcium<br />

Molecular and structural adaptations in islets such as increased mRNA of preproinsulin<br />

gene, reduction in islet size and increase in numbers, increase in islet cell proliferation<br />

Effects observed in post-weaning period<br />

Persistence of hyperinsulinemia, abnormal glucose tolerance test (GTT), leftward shift<br />

in the insulin secretory response to a glucose stimulus, increased gene expression of<br />

preproinsulin and related factors, hyperphagia, significant increases in body weight gain<br />

and adult-onset obesity<br />

Generational effects<br />

Female rats that underwent HC dietary modification spontaneously transmitted their<br />

phenotypes to their progeny<br />

Programming effects observed in HC progeny included:<br />

Fetal hyperinsulinemia<br />

Increased response by fetal islets to glucose and amino acids<br />

Increased preproinsulin gene expression in fetal islets<br />

Absence of hyperinsulinemia during suckling period and onset of hyperinsulinemia<br />

immediately after weaning<br />

Increased response to lower glucose concentrations by islets from 28-day-old progeny<br />

Increases in body weight gain in post-weaning period and full-blown obesity by postnatal<br />

day 100<br />

suckling period is an important phase of continued organogenesis, including<br />

pancreatic and brain development. 86<br />

The adaptation of the artificial rearing technique originally described by Hall 87<br />

has enabled our laboratory to investigate the immediate and long-term consequences<br />

of a high carbohydrate (HC) dietary modification during the suckling<br />

period in rats. 88 Newborn rat pups were raised away from their dams and given<br />

a high carbohydrate (HC) milk formula instead of the high fat rat milk received<br />

by naturally reared (mother-fed; MF) rat pups. 89,90 In order to establish that the


56 Oxidative Stress and Inflammatory Mechanisms<br />

artificial rearing technique per se did not cause any metabolic programming<br />

effects, newborn rat pups were also artificially reared on a high fat (HF) milk<br />

formula, the macronutrient composition of which was identical to that of rat<br />

milk. 90,91 Although MF, HF, and HC rat pups received a similar number of calories<br />

per day, the primary source of calories was switched from fats to carbohydrates<br />

for the HC rats. The mere change in the quality of nutrition during the suckling<br />

period (from fat-enriched rat milk to carbohydrate-enriched HC milk formula)<br />

resulted in metabolic programming due to the overlap of the HC dietary intervention<br />

and the critical post-natal period of pancreatic development in rats. 88,92,93<br />

The high carbohydrate nature of the HC formula (56% carbohydrate compared<br />

to 8% in rat milk and in the HF milk formula) increases insulin demand in neonatal<br />

rats. This altered nutritional environment induces compensatory adaptations in<br />

islet function to ensure that the immediate demand for insulin is met. However,<br />

permanent programming of these adaptations results in altered responses to nutritional<br />

experiences throughout life and adverse consequences in adulthood (Table<br />

4.2). 88,92,93<br />

METABOLIC PROGRAMMING IN HC RATS: IMMEDIATE EFFECTS<br />

Artificial rearing of rat pups on the HC milk formula resulted in the immediate<br />

onset (within the first 24 hours) of hyperinsulinemia when compared to MF and<br />

HF (experimental control) rats and persisted during the entire suckling period in<br />

the HC rats (Figure 4.2). 88,94 Interestingly, HC pups showed normal blood glucose<br />

levels during this period despite hyperinsulinemia, suggesting that the alterations<br />

in islet function are able to compensate for increased insulin demand to maintain<br />

0 22 4<br />

Birth<br />

12<br />

Hyperinsulinemia<br />

Cellular adaptations: changes in islet architecture<br />

Molecular adaptations: increases in preproinsulin<br />

gene transcription and related transcription<br />

factors<br />

Biochemical adaptations: increased response of<br />

isolated islets to various stimuli<br />

Hyperphagia and<br />

Significant increases<br />

in body weight and an<br />

upward trend in abnormal response to a<br />

body weight gain glucose load Full blown<br />

HC<br />

obesity<br />

Prenatal Suckling Post-weaning<br />

24 40 55 75 100<br />

Postnatal age (days)<br />

FIGURE 4.2 Overview of immediate and lasting consequences of HC dietary modification<br />

in the immediate post-natal period in first generation HC rats.


Metabolic Syndrome Due to Early Life Nutritional Modifications 57<br />

euglycemia. 94 In contrast, rats artificially reared on an HF formula (caloric distribution:<br />

carbohydrate 8%, protein 24%, and fat 68%) as an internal control for<br />

the artificial rearing technique were not hyperinsulinemic, suggesting that the<br />

metabolic programming of HC rats was the result of a change in the quality of<br />

nutrition and not a change in the mode of delivery of nutrition during the suckling<br />

period. 90<br />

In vitro studies of glucose-stimulated insulin secretion (GSIS) in isolated<br />

pancreatic islets showed that neonatal HC islets secreted significantly more insulin<br />

after both 10 and 60 min of glucose stimulation compared to MF islets at all<br />

glucose concentrations studied. 94 In association with this increased insulin secretory<br />

response to glucose, HC islets showed increases in (1) the enzyme activities<br />

of the low K m hexokinase, glyceraldehyde-3-phosphate dehydrogenase and pyruvate<br />

dehydrogenase complex and (2) increases in levels of glucose transporter<br />

protein-2 (Figure 4.2). 94 Up-regulation of glucose metabolism in HC islets suggests<br />

that a lowered glucose threshold for insulin secretion may, in part, be<br />

responsible for the development of hyperinsulinemia due to programmed hypersensitivity<br />

of pancreatic islets to glucose.<br />

In vivo insulin secretion is stimulated by numerous secretogogues including<br />

glucose, amino acids, fatty acids, and neuroendocrine and incretin factors. 95 While<br />

insulin secretion is primarily induced by glucose stimulation of the K ATP channeldependent<br />

pathway, acetylcholine, norepinephrine, glucagon, glucagon-like peptide-1,<br />

and other signals act on K ATP channel-dependent and -independent pathways<br />

as well as Ca 2+ channel-independent pathways to augment insulin secretion.<br />

95 Plasma glucagon-like peptide-1, an incretin hormone that promotes insulin<br />

secretion via increases in cAMP levels, 96 was increased in 12-day-old HC rats as<br />

were glucagon-like peptide-1 receptor mRNA levels in isolated HC islets. 97 Acetylcholine<br />

and glucagon-like peptide-1 potentiation of glucose-stimulated insulin<br />

secretion was greater in HC islets compared to MF islets, indicating increased<br />

responsiveness to these agonists. 97 Also, the activities of protein kinase C, protein<br />

kinase A, and calcium calmodulin kinase II were significantly higher in HC islets<br />

compared to MF islets, 86 as were the mRNA levels of adenylyl cyclase type VI. 97<br />

Furthermore, HC islets have reduced sensitivity to norepinephrine stimulation,<br />

which inhibits insulin secretion through activation of α-adrenergic receptors on<br />

β cells 98 compared to control islets. 97<br />

In addition to adaptations at the level of the insulin secretory process, molecular<br />

adaptations were observed in islets isolated from 12-day-old HC rats. The<br />

adaptations included significant increases in insulin biosynthesis and in the<br />

mRNA levels of the preproinsulin gene and other components of the cascade<br />

implicated in the regulation of the transcription of the preproinsulin gene in<br />

neonatal HC islets (Figure 4.2). 99 Gene array analysis of the mRNAs from neonatal<br />

HC islets indicated that several clusters of genes involved in a wide array<br />

of cellular functions (e.g., cell cycle regulation, protein synthesis, ion channels,<br />

and metabolic pathways) were up-regulated in these islets and may contribute to<br />

the onset of hyperinsulinemia in these rats. 100


58 Oxidative Stress and Inflammatory Mechanisms<br />

The overlap of the HC dietary modification with the period of post-natal<br />

development of the pancreas also resulted in several changes in the cellular<br />

architecture of neonatal HC islets (Table 4.2; Figure 4.2). These include: (1) a<br />

reduction in mean islet size and increase in the number of small-sized islets, (2)<br />

an increase in the number of islets per unit area, (3) greater apoptotic rates in<br />

islets compared to ductal epithelium, and (4) increases in islet cell replication<br />

as indicated by the presence of proliferating cell nuclear antigen in a large<br />

proportion of β cells in 12-day-old HC islets compared to islets from agematched<br />

MF rats. 101<br />

Metabolic programming effects were also observed in livers of neonatal HC<br />

rats and may be important for the establishment of the HC phenotype observed<br />

in adulthood. Precocious induction of the activities of glucokinase and malic<br />

enzyme were indicated by substantial increases in their activity levels in 10-dayold<br />

HC animals. 89 The lipogenic capacity and glycogen content were also<br />

increased in the livers of these rats. 89,102<br />

The hypothalamus is an important site for energy homeostasis and aberrations<br />

in the expression and function of a network of neuropeptides lead to either<br />

hyperphagia or hypophagia. 103,104 HC dietary modification resulted in significant<br />

increases in the protein content and gene expression of orexigenic signals and<br />

converse changes in the anorexigenic signals, suggesting that alterations leading<br />

to the eventual onset of obesity in the adulthood of these rats are evident very<br />

early in life (M.S. Patel, unpublished observations).<br />

PERSISTENT EFFECTS IN ADULTHOOD<br />

HC rats were hyperinsulinemic throughout the period of nutritional modification<br />

and continued to be hyperinsulinemic into adulthood even after weaning onto<br />

laboratory rodent diets on post-natal day 24 (Figure 4.2). 90,91 Studies of 100-dayold<br />

adult rats showed that HC animals had significantly higher plasma insulin<br />

levels compared with age-matched MF rats. 105 In addition to persistent hyperinsulinemia,<br />

adult HC rats showed impaired glucose tolerance compared with<br />

controls, even though they were euglycemic, suggesting a state of insulin resistance<br />

(Figure 4.3). 91<br />

As in the case of 12-day-old HF rats, adult HF rats were not hyperinsulinemic<br />

and did not show altered growth or food intake compared to MF rats. 90 While<br />

HC and MF rats showed similar body weights during the suckling period, 94 HC<br />

rats experienced increased growth rates after weaning in association with hyperphagia<br />

that resulted in full-blown obesity by day 100 (Figure 4.2). 90,106<br />

In addition, islets from adult HC rats showed increases in GSIS, low K m<br />

hexokinase activity, and Ca 2+ -independent insulin release similar to results from<br />

12-day-old HC islets. 105 Additionally, increases in insulin producing cell masses<br />

in adult HC pancreas were noted. 91 At the molecular level, increased preproinsulin<br />

gene transcription with concomitant increases in mRNA levels of transcription<br />

factors regulating its gene expression and increases in the expression of several<br />

clusters of genes as indicated by gene array analysis 105 were observed in


Metabolic Syndrome Due to Early Life Nutritional Modifications 59<br />

Plasma glucose (mM)<br />

Plasma glucose (mM) Plasma glucose (mM)<br />

16<br />

12<br />

8<br />

4<br />

20<br />

16<br />

12<br />

8<br />

24<br />

20<br />

16<br />

12<br />

8<br />

FIGURE 4.3 Plasma glucose and insulin levels in HC male rats after an oral glucose<br />

tolerance load (2 g/kg body weight) on post-natal days 64, 78, and 270. 91 Completely<br />

overlapping points for plasma glucose of 270-day old MF and HC rats are shown with<br />

open circles only.<br />

100-day-old HC islets, suggesting that the molecular effects observed in neonatal<br />

islets persist into adulthood. Adult onset obesity was accompanied by increases<br />

in the lipogenic capacities of the liver and epididymal adipose tissue, increases<br />

in the epididymal adipose tissue weight, and marked increases in the cell sizes<br />

of epididymal and omental adipose tissues of 100-day-old male HC rats. 90<br />

GENERATIONAL EFFECTS<br />

Day 64<br />

0<br />

0 20 40 60 80 100 120 0 20 40 60 80 100 120<br />

Day 78<br />

Day 270<br />

A notable observation from the HC model is that the progeny of HC female rats<br />

that underwent the HC dietary modification in their immediate post-natal life<br />

Plasma insulin (pM)<br />

Plasma insulin (pM)<br />

800<br />

600<br />

400<br />

200<br />

1600<br />

1200<br />

MF<br />

HC<br />

0<br />

0 20 40 60 80 100 120 0 20 40 60 80 100 120<br />

Plasma insulin (pM)<br />

800<br />

400<br />

1600<br />

1200<br />

0 20 40 60 80 100 120 0 20 40 60 80 100 120<br />

Time (min after glucose bolus) Time (min after glucose bolus)<br />

800<br />

400


60 Oxidative Stress and Inflammatory Mechanisms<br />

Term fetus<br />

Hyperinsulinemia, normoglycemia<br />

Increased insulin secretory response<br />

by islets<br />

Increased expression of preproinsulin<br />

gene<br />

Maternal<br />

hyperinsulinemia,<br />

increased body<br />

weight gain during<br />

pregnancy<br />

Prenatal Suckling Post-weaning<br />

0 210 12 24 28 55 75 10<br />

Term fetus<br />

Hyperinsulinemia<br />

Increased sensitivity to basal glucose<br />

Increased expression of preproinsulin<br />

gene<br />

Increases in body<br />

weight gain<br />

Postnatal age (days)<br />

Full blown<br />

obesity<br />

FIGURE 4.4 Overview of generational effects in progeny (second generation HC rats)<br />

resulting from HC dietary modifications in female rats.<br />

spontaneously acquired the HC phenotype (chronic hyperinsulinemia and adult<br />

onset obesity) without undergoing dietary modifications (Figure 4.4). 106 The<br />

intrauterine environment in the HC female rat was characterized by hyperinsulinemia<br />

and normoglycemia. 106–108 We recently determined that the HC female rats<br />

consumed significantly increased amounts of food and gained significantly more<br />

weight during gestation compared to pregnant MF controls; they were hyperleptinemic<br />

on gestational day 21 (M.S Patel, unpublished observations).<br />

HC term fetuses (gestational day 21) were hyperinsulinemic and normoglycemic.<br />

Fetal hyperinsulinemia was accompanied by increased pancreatic insulin<br />

content, increased gene expression of preproinsulin and pancreatic duodenal<br />

transcription factor-1, and increased insulin secretory responses to various secretogogues<br />

(Figure 4.4). 108 Although no significant changes were noted in plasma<br />

insulin levels during the suckling period between the second generation HC (2-<br />

HC) and MF rats immediately after weaning, the plasma insulin levels of 2-HC<br />

rats were significantly increased (first observed on post-natal day 26; Figure<br />

4.4). 107 As observed in term fetuses, there was a leftward shift in the insulin<br />

secretory response to a glucose stimulus and increased preproinsulin gene transcription<br />

in islets from 28-day old 2-HC rats. 107<br />

Similar to the observations in first generation HC rats, no significant differences<br />

were noted in the body weights of 2-HC rats compared to age-matched<br />

MF rats up to post-natal day 55. After post-natal day 55, significant increases in<br />

body weight were observed with full blown obesity evident on approximately<br />

post-natal day 100. 106 This correlates with significant increases in the weight of<br />

the adipose tissues, increases in the size of the adipocytes, and increases in the<br />

activities of the lipogenic enzymes (fatty acid synthase and glucose-6-phosphate<br />

dehydrogenase) in both liver and adipose tissue of 100-day old 2-HC rats. 106


Metabolic Syndrome Due to Early Life Nutritional Modifications 61<br />

A state of insulin resistance was indicated by decreased content of glycogen<br />

in livers and muscles of 100-day-old 2-HC male rats. 109 The decrease was associated<br />

with a decrease in the enzyme activity of glycogen synthase and its putative<br />

upstream activators. 109 In contrast, the situation was reversed in the epididymal<br />

adipose tissues of these rats. 110<br />

Are the metabolic programming effects induced by early life HC dietary<br />

modifications in rat pups reversible phenomena? In order to address this question,<br />

we pair-fed HC female rats from the time of weaning to the amount of feed<br />

consumed by age-matched MF female rats on a daily basis. Interestingly, such a<br />

dietary control in HC female rats resulted in reductions of their plasma insulin<br />

levels and body weight gains to the levels observed in age-matched female rats. 108<br />

During pregnancy, the plasma insulin levels and body weights were similar in<br />

the pair-fed HC and age-matched MF female rats. 108 Such improvements in<br />

pregnancy conditions in the pair-fed HC female rats prevented the transmission<br />

of the HC phenotype to their progeny. 108 These results indicate that dietary<br />

regulation in HC rats results in the reversal of the metabolic programming effects<br />

with good prognosis both for rats of the same generation and the subsequent<br />

generation.<br />

The mechanisms responsible for metabolic programming due to altered nutritional<br />

experiences in early life are not well understood. Malorganization (functional<br />

and/or structural) of target organs has been suggested as a plausible mechanism.<br />

111 In animal models for pre-natal metabolic programming (protein<br />

malnourishment, gestational diabetes, and caloric restriction), such adaptations<br />

in target organs were reported. 30 In the HC rat model, we observed functional<br />

and/or structural changes in neonatal HC rats in target organs such as the islets,<br />

gut, hypothalamus, and liver. All of these organs are potentially important for<br />

maintaining glucose homeostasis. As indicated in Figure 4.5, it is possible that<br />

adaptations in the these organs and the resultant cross-talks among them are<br />

necessary for survival of these rat pups in the face of the HC dietary modification<br />

but since they occur during the period of immediate post-natal development, such<br />

adaptations become permanent and in the post-weaning period, predispose these<br />

animals to adult onset obesity and concomitant health issues.<br />

Various interactions between target organs are suggested in Figure 4.5. The<br />

HC milk formula may induce alterations in stomach- and gut-derived factors<br />

which, through specific receptor-mediated signaling, may alter the functional<br />

capabilities of organs such as the islets and the hypothalamus. Similarly, insulin<br />

via receptors present in the islets and the hypothalamus may exert altered autocrine<br />

and endocrine functions to support the HC phenotypes in these rats. Several<br />

neuropeptides enumerated in the hypothalamus have principal roles in the energy<br />

homeostasis of the body, insulin secretion by islets, insulin sensitivity of target<br />

organs, etc. Therefore, a multifactorial mechanism may be responsible for the<br />

onset and persistence of hyperinsulinemia in HC rats which predisposes these<br />

rats for insulin resistance leading to altered glucose tolerance and obesity in<br />

adulthood.


62 Oxidative Stress and Inflammatory Mechanisms<br />

Brain (hypothalamus)<br />

Alterations in energy<br />

homeostasisrelated<br />

neuropeptides<br />

Diet<br />

(high carbohydrate)<br />

Pancreas<br />

(islets)<br />

Hyperinsulinemia<br />

(immediate response)<br />

Stomach and gut<br />

(gut-related factors)<br />

Metabolic adaptations in peripheral tissues<br />

Persistence of hyperinsulinemia<br />

Obesity and other complications in adulthood<br />

Unfavorable intrauterine environment in female rats<br />

Transmission of the maternal phenotype to the progeny<br />

(perpetuation of the metabolic programming effects)<br />

FIGURE 4.5 Postulated scheme leading to development of the HC metabolic phenotype<br />

in adulthood due to high-carbohydrate dietary modifications in neonatal rats. The direct<br />

effects of the dietary treatment on target organs and possible cross-talks among these<br />

organs leading to the adult phenotype are indicated.<br />

CORRELATION OF ALTERED NUTRITIONAL EXPERIENCE IN<br />

EARLY LIFE TO SUBSEQUENT HIGH INCIDENCE OF OBESITY<br />

<strong>AND</strong> METABOLIC SYNDROME<br />

As indicated, nutritional experiences in utero and/or in the immediate post-natal<br />

period (infancy) can to a significant degree influence adult metabolic phenotype<br />

as these are periods of rapid development of the organism. Because the main goal<br />

is survival, an organism adapts to the altered environment by necessary alterations<br />

in target tissues that help it “tide over” the situation but such adaptations lead to<br />

unfavorable situations later in life. Both epidemiological data and results from<br />

animal studies indicate the importance of adequate nutrition during pregnancy.<br />

Studies of the long-term consequences of an altered nutritional experience in<br />

early post-natal life indicate the importance of this phase of life for metabolic<br />

programming effects. Dietary habits for all ages have undergone tremendous<br />

changes over the past several decades. The present obesity epidemic, to a large<br />

measure, may be the result of such changes. Extrapolation of data obtained from<br />

HC rat models suggests that post-natal increased consumption of carbohydrates


Metabolic Syndrome Due to Early Life Nutritional Modifications 63<br />

by infants (formula feeding with early introduction of carbohydrate-rich supplements<br />

such as cereals, fruits juices, etc.) in Western societies may be partly<br />

responsible for the increase in the incidence of obesity. This effect is exacerbated<br />

by the mode of feeding (bottle, spoon, etc., resulting in overfeeding). Supplementation<br />

of milk (breast or formula) with early introduction of carbohydrateenriched<br />

baby foods and overfeeding may result in malprogramming effects in<br />

these babies, leading to adult onset obesity and attendant diseases as observed in<br />

the HC rat model.<br />

The spontaneous transfer of the HC phenotype to the progeny and results<br />

from our studies of maternal high-fat diet-induced effects in the next generation<br />

indicate that maternal obesity primes obesity in progeny. Our results suggest that<br />

females that are overnourished or exposed to increased carbohydrate intake in<br />

infancy may be obese and insulin resistant during pregnancy and, due to an<br />

unfavorable intrauterine environment, are at increased risk for the establishment<br />

of a vicious cycle of transmission of their metabolic phenotype to their progeny.<br />

The alarming increase in the numbers of overweight and obese individuals<br />

in the United States suggests that a significant number of pregnancies will not<br />

be under conditions of optimal health due to increased body weight, moderate<br />

hyperinsulinemia, and mild insulin resistance in these females. Fetal development<br />

under such conditions may predispose the progeny for the development of obesity<br />

in adulthood. This generational effect may be amplified by dietary habits in<br />

infancy and lifestyle later in adulthood such that the maternal intrauterine environment<br />

becomes more and more favorable for metabolic malprogramming of<br />

the progeny from one generation to the next.<br />

ACKNOWLEDGMENT<br />

This work was supported in part by National Institute of Diabetes and Digestive<br />

and Kidney Diseases grant DK-61518.<br />

REFERENCES<br />

1. Daniels, J. Obesity: America’s epidemic, Am. J. Nurs. 106, 40, 2006.<br />

2. Finkelstein, E.A., Ruhm, C.J., and Kosa, K.M. Economic causes and consequences<br />

of obesity, Annu. Rev. Public Health 26, 239, 2005.<br />

3. Flegal, K.M. et al. Prevalence and trends in obesity among U.S. adults, 1999–2000,<br />

JAMA 288, 1723, 2002.<br />

4. Flegal, K.M. Epidemiologic aspects of overweight and obesity in the United<br />

States, Physiol. Behav. 86, 599, 2005.<br />

5. Haffner, S. and Taegtmeyer, H. Epidemic obesity and the metabolic syndrome,<br />

Circulation 108, 1541, 2003.<br />

6. Zimmet, P., Alberti, K.G., and Shaw, J. Global and societal implications of the<br />

diabetes epidemic, Nature 414, 782, 2001.<br />

7. Stein, C.J. and Colditz, G.A. The epidemic of obesity, J. Clin. Endocrinol. Metab.<br />

89, 2522, 2004.


64 Oxidative Stress and Inflammatory Mechanisms<br />

8. Mokdad, A.H. et al. Prevalence of obesity, diabetes, and obesity-related health<br />

risk factors, 2001, JAMA 289, 76, 2003.<br />

9. Reaven, G.M. Banting Lecture, 1988: role of insulin resistance in human disease,<br />

Diabetes 37, 1595, 1988.<br />

10. Armitage, J.A. et al. Developmental programming of the metabolic syndrome by<br />

maternal nutritional imbalance: how strong is the evidence from experimental<br />

models in mammals? J. Physiol. 561, 355, 2004.<br />

11. Schwimmer, J.B. et al. Obesity, insulin resistance, and other clinicopathological<br />

correlates of pediatric nonalcoholic fatty liver disease, J. Pediatr. 143, 500, 2003.<br />

12. Csabi, G. et al. Presence of metabolic cardiovascular syndrome in obese children,<br />

Eur. J. Pediatr. 159, 91, 2000.<br />

13. Slyper, A.H. The pediatric obesity epidemic: causes and controversies, J. Clin.<br />

Endocrinol. Metab. 89, 2540, 2004.<br />

14. Molnar, D. The prevalence of the metabolic syndrome and type 2 diabetes mellitus<br />

in children and adolescents, Int. J. Obes. Relat. Metab. Disord. 28, Suppl. 3, S70,<br />

2004.<br />

15. Heindel, J.J. Endocrine disruptors and the obesity epidemic, Toxicol. Sci. 76, 247,<br />

2003.<br />

16. Trends in intake of energy and macronutrients, United States, 1971–2000, Morb.<br />

Mortal. Wkly. Rep. 53, 80, 2004.<br />

17. Lucas, A. Programming by early nutrition in man, Ciba Found. Symp. 156, 38,<br />

1991.<br />

18. Barker, D.J. et al. Fetal nutrition and cardiovascular disease in adult life, Lancet<br />

341, 938, 1993.<br />

19. Barker, D.J. Fetal origins of coronary heart disease, Brit. Med. J. 311, 171, 1995.<br />

20. Hales, C.N. and Barker, D.J. Type 2 (non-insulin-dependent) diabetes mellitus:<br />

the thrifty phenotype hypothesis, Diabetologia 35, 595, 1992.<br />

21. Prentice, A.M. Early influences on human energy regulation: thrifty genotypes<br />

and thrifty phenotypes, Physiol. Behav. 86, 640, 2005.<br />

22. Holness, M.J., Langdown, M.L., and Sugden, M.C. Early-life programming of<br />

susceptibility to dysregulation of glucose metabolism and the development of type<br />

2 diabetes mellitus, Biochem. J. 349, Pt 3, 657, 2000.<br />

23. Fowden, A.L., Giussani, D.A., and Forhead, A.J. Intrauterine programming of<br />

physiological systems: causes and consequences, Physiology 21, 29, 2006.<br />

24. Hales, C.N. and Barker, D.J. Thrifty phenotype hypothesis, Br. Med. Bull. 60, 5,<br />

2001.<br />

25. Neel, J.V. Diabetes mellitus: a “thrifty” genotype rendered detrimental by<br />

“progress”? Am. J. Hum. Genet. 14, 353, 1962.<br />

26. Barker, D.J. et al. Weight in infancy and death from ischaemic heart disease,<br />

Lancet 2, 577, 1989.<br />

27. Godfrey, K.M. et al. Maternal birthweight and diet in pregnancy in relation to the<br />

infant’s thinness at birth, Br. J. Obstet. Gynaecol. 104, 663, 1997,<br />

28. Barker, D.J. Fetal nutrition and cardiovascular disease in later life, Br. Med. Bull.<br />

53, 96, 1997.<br />

29. Petry, C.J. and Hales, C.N. Intrauterine Development and Its Relationship with<br />

Type 2 Diabetes Mellitus, John Wiley & Sons, Chichester, 1999.<br />

30. Petry, C.J., Ozanne, S.E., and Hales, C.N. Programming of intermediary metabolism,<br />

Mol. Cell. Endocrinol. 185, 81, 2001.


Metabolic Syndrome Due to Early Life Nutritional Modifications 65<br />

31. Betram, C.E. and Hanson, M.A. Animal models and programming of the metabolic<br />

syndrome, Br. Med. Bull. 60, 103, 2001.<br />

32. Van Assche, F.A., Holemans, K., and Aerts, L. Long-term consequences for<br />

offspring of diabetes during pregnancy, Br. Med. Bull. 60, 173, 2001.<br />

33. Holemans, K., Aerts, L., and Van Assche, F.A. Lifetime consequences of abnormal<br />

fetal pancreatic development, J. Physiol. 547, 11, 2003.<br />

34. Snoeck, A. et al. Effect of a low protein diet during pregnancy on the fetal rat<br />

endocrine pancreas, Biol. Neonate 57, 107, 1990.<br />

35. Dahri, S. et al. Islet function in offspring of mothers on low-protein diet during<br />

gestation, Diabetes 40, Suppl. 2, 115, 1991.<br />

36. Plagemann, A. et al. Hypothalamic nuclei are malformed in weanling offspring<br />

of low protein malnourished rat dams, J. Nutr. 130, 2582, 2000.<br />

37. Berney, D.M. et al. The effects of maternal protein deprivation on the fetal rat<br />

pancreas: major structural changes and their recuperation, J. Pathol. 183, 109,<br />

1997.<br />

38. Latorraca, M.Q. et al. Protein deficiency and nutritional recovery modulate insulin<br />

secretion and the early steps of insulin action in rats, J. Nutr. 128, 1643, 1998.<br />

39. Ozanne, S.E. and Hales, C.N. The long-term consequences of intra-uterine protein<br />

malnutrition for glucose metabolism, Proc. Nutr. Soc. 58, 615, 1999.<br />

40. Bennis-Taleb, N. et al. A low-protein isocaloric diet during gestation affects brain<br />

development and alters permanently cerebral cortex blood vessels in rat offspring,<br />

J. Nutr. 129, 1613, 1999.<br />

41. Desai, M. et al. Organ-selective growth in the offspring of protein-restricted<br />

mothers, Br. J. Nutr. 76, 591, 1996.<br />

42. Malandro, M.S. et al. Effect of low-protein diet-induced intrauterine growth retardation<br />

on rat placental amino acid transport, Am. J. Physiol. 271, C295, 1996.<br />

43. Rees, W.D. et al. The effects of maternal protein restriction on the growth of the<br />

rat fetus and its amino acid supply, Br. J. Nutr. 81, 243, 1999.<br />

44. Hoet, J.J. and Hanson, M.A. Intrauterine nutrition: its importance during critical<br />

periods for cardiovascular and endocrine development, J. Physiol. 514, Pt. 3, 617,<br />

1999.<br />

45. Petrik, J. et al. A low protein diet alters the balance of islet cell replication and<br />

apoptosis in the fetal and neonatal rat and is associated with a reduced pancreatic<br />

expression of insulin-like growth factor-II, Endocrinology 140, 4861, 1999.<br />

46. Cherif, H. et al. Stimulatory effects of taurine on insulin secretion by fetal rat<br />

islets cultured in vitro, J. Endocrinol. 151, 501, 1996.<br />

47. Langley-Evans, S.C. Hypertension induced by foetal exposure to a maternal lowprotein<br />

diet in the rat is prevented by pharmacological blockade of maternal<br />

glucocorticoid synthesis, J. Hypertens. 15, 537, 1997.<br />

48. Perkins, M.N. et al. Activation of brown adipose tissue thermogenesis by the<br />

ventromedial hypothalamus, Nature 289, 401, 1981.<br />

49. Krukoff, T.L. et al. Expression of c-fos protein in rat brain elicited by electrical<br />

and chemical stimulation of the hypothalamic paraventricular nucleus, Neuroendocrinology<br />

59, 590, 1994.<br />

50. Garofano, A., Czernichow, P., and Breant, B. Beta-cell mass and proliferation<br />

following late fetal and early post-natal malnutrition in the rat, Diabetologia 41,<br />

1114, 1998.


66 Oxidative Stress and Inflammatory Mechanisms<br />

51. Garofano, A., Czernichow, P., and Breant, B. Effect of ageing on beta-cell mass<br />

and function in rats malnourished during the perinatal period, Diabetologia 42,<br />

711, 1999.<br />

52. Aerts, L., Holemans, K., and Van Assche, F.A. Maternal diabetes during pregnancy:<br />

consequences for the offspring, Diabetes Metab. Rev. 6, 147, 1990.<br />

53. Reusens-Billen, B. et al. Cell proliferation in pancreatic islets of rat fetuses and<br />

neonates from normal and diabetic mothers: an in vitro and in vivo study, Horm.<br />

Metab. Res. 16, 565, 1984.<br />

54. Eriksson, U. et al. Diabetes in pregnancy: effects on the foetal and newborn rat<br />

with particular regard to body weight, serum insulin concentration and pancreatic<br />

contents of insulin, glucagon and somatostatin, Acta Endocrinol. 94, 354, 1980.<br />

55. Kervran, A., Guillaume, M., and Jost, A. The endocrine pancreas of the fetus from<br />

diabetic pregnant rat, Diabetologia 15, 387, 1978.<br />

56. Bihoreau, M.T., Ktorza, A., and Picon, L. Gestational hyperglycaemia and insulin<br />

release by the fetal rat pancreas in vitro: effect of amino acids and glyceraldehyde,<br />

Diabetologia 29, 434, 1986.<br />

57. Plagemann, A. et al. Perinatal elevation of hypothalamic insulin, acquired malformation<br />

of hypothalamic galaninergic neurons, and syndrome x-like alterations<br />

in adulthood of neonatally overfed rats, Brain Res. 836, 146, 1999.<br />

58. Gallou-Kabani, C. and Junien, C. Nutritional epigenomics of metabolic syndrome:<br />

new perspective against the epidemic, Diabetes 54, 1899, 2005.<br />

59. Srinivasan, M. et al. Maternal high fat diet consumption results in fetal malprogramming<br />

predisposing to the onset of metabolic syndrome-like phenotype in<br />

their adulthood, Am. J. Physiol. Endocrinol. Metab. 2006.<br />

60. Cerf, M.E. et al. Islet cell response in the neonatal rat after exposure to a highfat<br />

diet during pregnancy, Am. J. Physiol. Regul. Integr. Comp. Physiol. 288,<br />

R1122, 2005.<br />

61. Guo, F. and Jen, K.L. High-fat feeding during pregnancy and lactation affects<br />

offspring metabolism in rats, Physiol. Behav. 57, 681, 1995.<br />

62. Taylor, P.D. et al. Impaired glucose homeostasis and mitochondrial abnormalities<br />

in offspring of rats fed a fat-rich diet in pregnancy, Am. J. Physiol. Regul. Integr.<br />

Comp. Physiol. 288, R134, 2005.<br />

63. Khan, I.Y. et al. A high-fat diet during rat pregnancy or suckling induces cardiovascular<br />

dysfunction in adult offspring, Am. J. Physiol. Regul. Integr. Comp.<br />

Physiol. 288, R127, 2005.<br />

64. Khan, I.Y. et al. Gender-linked hypertension in offspring of lard-fed pregnant rats,<br />

Hypertension 41, 168, 2003.<br />

65. Armitage, J.A., Taylor, P.D., Poston, L., Experimental models of developmental<br />

programming: consequences of exposure to an energy rich diet during development,<br />

J. Physiol. 565, 3, 2005.<br />

66. Palinski, W. et al. Maternal hypercholesterolemia and treatment during pregnancy<br />

influence the long-term progression of atherosclerosis in offspring of rabbits, Circ.<br />

Res. 89, 991, 2001.<br />

67. Kozak, R., Richy, S., and Beck, B., Persistent alterations in neuropeptide Y release<br />

in the paraventricular nucleus of rats subjected to dietary manipulation during<br />

early life, Eur. J. Neurosci. 21, 2887, 2005.<br />

68. Mortensen, E.L. et al. The association between duration of breastfeeding and adult<br />

intelligence, JAMA 287, 2365, 2002,


Metabolic Syndrome Due to Early Life Nutritional Modifications 67<br />

69. Haisma, H. et al. Breast milk and energy intake in exclusively, predominantly,<br />

and partially breast-fed infants, Eur. J. Clin. Nutr. 57, 1633, 2003.<br />

70. Baker, J.L. et al. Maternal prepregnant body mass index, duration of breastfeeding,<br />

and timing of complementary food introduction are associated with infant weight<br />

gain, Am. J. Clin. Nutr. 80, 1579, 2004.<br />

71. Owen, C.G. et al. The effect of breastfeeding on mean body mass index throughout<br />

life: a quantitative review of published and unpublished observational evidence,<br />

Am. J. Clin. Nutr. 82, 1298, 2005.<br />

72. Owen, C.G. et al. Effect of infant feeding on the risk of obesity across the life<br />

course: a quantitative review of published evidence, Pediatrics 115, 1367, 2005.<br />

73. Armstrong, J. and Reilly, J.J. Breastfeeding and lowering the risk of childhood<br />

obesity, Lancet 359, 2003, 2002.<br />

74. Heinig, M.J. et al. Energy and protein intakes of breast-fed and formula-fed infants<br />

during the first year of life and their association with growth velocity: the DAR-<br />

LING Study, Am. J. Clin. Nutr. 58, 152, 1993.<br />

75. Lucas, A. et al. Breast versus bottle: endocrine responses are different with formula<br />

feeding, Lancet 1, 1267, 1980.<br />

76. McCance, R.A. Food, growth, and time, Lancet 2, 671, 1962.<br />

77. Faust, I.M., Johnson, P.R., and Hirsch, J. Long-term effects of early nutritional<br />

experience on the development of obesity in the rat, J. Nutr. 110, 2027, 1980,<br />

78. Plagemann, A. et al. Obesity and enhanced diabetes and cardiovascular risk in<br />

adult rats due to early post-natal overfeeding, Exp. Clin. Endocrinol. 99, 154, 1992.<br />

79. Waterland, R.A. and Garza, C. Early post-natal nutrition determines adult pancreatic<br />

glucose-responsive insulin secretion and islet gene expression in rats, J.<br />

Nutr. 132, 357, 2002.<br />

80. Davidowa, H. and Plagemann, A. Hypothalamic neurons of post-natally overfed,<br />

overweight rats respond differentially to corticotropin-releasing hormones, Neurosci.<br />

Lett. 371, 64, 2004.<br />

81. Plagemann, A. et al. Elevation of hypothalamic neuropeptide Y-neurons in adult<br />

offspring of diabetic mother rats, Neuroreport 10, 3211, 1999.<br />

82. Plagemann, A. et al. Increased number of galanin-neurons in the paraventricular<br />

hypothalamic nucleus of neonatally overfed weanling rats, Brain Res. 818, 160,<br />

1999.<br />

83. Walker, C.D. Nutritional aspects modulating brain development and the responses<br />

to stress in early neonatal life, Progr. Neuropsychopharmacol. Biol. Psychiatr. 29,<br />

1249, 2005.<br />

84. Li, Y., Plagemann, A., and Davidowa, H. Increased inhibition by agouti-related<br />

peptide of ventromedial hypothalamic neurons in rats overweight due to early<br />

post-natal overfeeding, Neurosci. Lett. 330, 33, 2002.<br />

85. Cryer, A. and Jones, H.M. The development of white adipose tissue: effect of<br />

litter size on the lipoprotein lipase activity of four adipose-tissue depots, serum<br />

immunoreactive insulin and tissue cellularity during the first year of life in male<br />

and female rats, Biochem. J. 186, 805, 1980.<br />

86. Kaung, H.L. Growth dynamics of pancreatic islet cell populations during fetal<br />

and neonatal development of the rat, Dev. Dyn. 200, 163, 1994.<br />

87. Hall, W.G. Weaning and growth of artificially reared rats, Science 190, 1313, 1975.<br />

88. Srinivasan, M. et al. Neonatal nutrition: metabolic programming of pancreatic<br />

islets and obesity, Exp. Biol. Med. 228, 15, 2003.


68 Oxidative Stress and Inflammatory Mechanisms<br />

89. Haney, P.M. et al. Precocious induction of hepatic glucokinase and malic enzyme<br />

in artificially reared rat pups fed a high-carbohydrate diet, Arch. Biochem. Biophys.<br />

244, 787, 1986.<br />

90. Hiremagular, B.K., Johanning, G.L., and Patel, M.S. Long-term effects of feeding<br />

of high carbohydrate diet in preweaning by gastrosomy: a new rat model for<br />

obesity, Int. J. Obes. 17, 495, 1993.<br />

91. Vadlamudi, S. et al. Long-term effects on pancreatic function of feeding a HC<br />

formula to rats during the preweaning period, Am. J. Physiol. 265, E565, 1993.<br />

92. Patel, M.S. and Srinivasan, M. Metabolic programming: causes and consequences,<br />

J. Biol. Chem. 277, 1629, 2002.<br />

93. Patel, M.S. and Srinivasan, M. Metabolic Programming as a Consequence of the<br />

Nutritional Environment during Fetal and the Immediate Post-natal Periods, Cambridge<br />

University Press, New York, 2006.<br />

94. Aalinkeel, R. et al. A dietary intervention (high carbohydrate) during the neonatal<br />

period causes islet dysfunction in rats, Am. J. Physiol. 277, E1061, 1999.<br />

95. Bratanova-Tochkova, T.K. et al. Triggering and augmentation mechanisms, granule<br />

pools, and biphasic insulin secretion, Diabetes 51, Suppl. 1, S83, 2002.<br />

96. Kieffer, T.J. and Habener, J.F. The glucagon-like peptides, Endocr. Rev. 20, 876,<br />

1999.<br />

97. Srinivasan, M. et al. Adaptive changes in insulin secretion by islets from neonatal<br />

rats raised on a high-carbohydrate formula, Am. J. Physiol. Endocrinol. Metab.<br />

279, E1347, 2000.<br />

98. Sharp, G.W. Mechanisms of inhibition of insulin release, Am. J. Physiol. 271,<br />

C1781, 1996.<br />

99. Srinivasan, M. et al. Molecular adaptations in islets from neonatal rats reared<br />

artificially on a high carbohydrate milk formula, J. Nutr. Biochem. 12, 575, 2001.<br />

100. Song, F. et al. Use of a cDNA array for the identification of genes induced in<br />

islets of suckling rats by a high-carbohydrate nutritional intervention, Diabetes<br />

50, 2053, 2001.<br />

101. Petrik, J. et al. A long-term high-carbohydrate diet causes an altered ontogeny of<br />

pancreatic islets of Langerhans in the neonatal rat, Pediatr. Res. 49, 84, 2001.<br />

102. Hiremagalur, B.K., Johanning, G.L., and Patel, M.S. Alterations in hepatic lipogenic<br />

capacity in rat pups artificially reared on a milk-substitute formula high in<br />

carbohydrate or medium-chain triglycerides, J. Nutr. Biochem. 3, 474, 1992.<br />

103. Schwartz, M.W. et al. Central nervous system control of food intake, Nature 404,<br />

661, 2000.<br />

104. Seeley, R.J. and Woods, S.C. Monitoring of stored and available fuel by the CNS:<br />

implications for obesity, Nat. Rev. Neurosci. 4, 901, 2003.<br />

105. Aalinkeel, R. et al. Programming into adulthood of islet adaptations induced by<br />

early nutritional intervention in the rat, Am. J. Physiol. Endocrinol. Metab. 281,<br />

E640, 2001.<br />

106. Vadlamudi, S., Kalhan, S.C., and Patel, M.S. Persistence of metabolic consequences<br />

in the progeny of rats fed a HC formula in their early post-natal life, Am.<br />

J. Physiol. 269, E731, 1995.<br />

107. Srinivasan, M. et al. Programming of islet functions in the progeny of hyperinsulinemic/obese<br />

rats, Diabetes 52, 984, 2003.<br />

108. Srinivasan, M. et al. Maternal hyperinsulinemia predisposes rat fetuses for hyperinsulinemia,<br />

and adult-onset obesity and maternal mild food restriction reverses<br />

this phenotype, Am. J. Physiol. Endocrinol. Metab. 290, E129, 2006.


Metabolic Syndrome Due to Early Life Nutritional Modifications 69<br />

109. Srinivasan, M., Vadlamudi, S., and Patel, M.S. Glycogen synthase regulation in<br />

hyperinsulinemic/obese progeny of rats fed a high carbohydrate formula in their<br />

infancy, Int. J. Obes. Relat. Metab. Disord. 20, 981, 1996.<br />

110. Srinivasan, M. et al. Glycogen synthase activation in the epididymal adipose tissue<br />

from chronic hyperinsulinemic/obese rats, Nutr. Biochem. 9, 81, 1998.<br />

111. Waterland, R.A. and Garza, C. Potential mechanisms of metabolic imprinting that<br />

lead to chronic disease, Am. J. Clin. Nutr. 69, 179, 1999.


5<br />

CONTENTS<br />

Oxidative Stress and<br />

Antioxidants in the<br />

Perinatal Period<br />

Hiromichi Shoji, Yuichiro Yamashiro, and<br />

Berthold Koletzko<br />

Introduction .........................................................................................................71<br />

Oxidative Stress ..................................................................................................72<br />

Pregnancy: Oxidative Stress ...............................................................................74<br />

Preeclampsia..............................................................................................75<br />

Oxidative Stress at Childbirth.............................................................................75<br />

Perinatal Asphyxia.....................................................................................76<br />

Use of 100% Oxygen in Resuscitation.....................................................77<br />

Oxidative Stress in Premature Infants ................................................................78<br />

Bronchopulmonary Dysplasia ...................................................................79<br />

Neonatal Necrotizing Enterocolitis ...........................................................80<br />

Periventricular Leukomalacia....................................................................80<br />

Retinopathy of Prematurity .......................................................................81<br />

Human Milk and Antioxidative Protection.........................................................82<br />

Conclusions .........................................................................................................84<br />

Acknowledgments ...............................................................................................84<br />

References ...........................................................................................................85<br />

INTRODUCTION<br />

Oxidative stress is thought to be implicated in many pathologic processes of<br />

human disorders. 1 Pregnancy is a physiological state of oxidative stress accompanied<br />

by high metabolic demands and elevated requirements for tissue oxygen. 2<br />

Childbirth is also accompanied by increased oxidative stress. Numerous disorders<br />

in infancy have been linked with oxidative stress, while the role of oxidative<br />

stress in the pathogenesis and progression of these diseases is only partially<br />

defined. 3 In 1988, Saugstad advocated “oxygen radical disease in neonatology”<br />

as a unifying hypothesis for the role of oxidative stress in a wide range of neonatal<br />

71


72 Oxidative Stress and Inflammatory Mechanisms<br />

morbidities, implying different manifestations of the same condition. 4 Thus it has<br />

become important for clinicians who treat patients in the perinatal period to<br />

understand oxidative stress. The purpose of this chapter is to review information<br />

about the role of oxidative stress in pregnancy and post-partum.<br />

<strong>OXIDATIVE</strong> <strong>STRESS</strong><br />

Free radicals are highly reactive chemical molecules containing one or more<br />

unpaired electrons. They donate or take electrons from other molecules in an<br />

attempt to pair their elections and generate a more stable species. Reactive oxygen<br />

species (ROS) is a collective term that includes both the oxygen free radicals (the<br />

O 2· – superoxide and the OH· hydroxyl radical) and some non-radical (hydrogen<br />

peroxide, H 2O 2) derivatives of oxygen. ROS are normally produced in living<br />

organisms with the potential of reacting with almost all types of molecules in<br />

living cells. ROS may be generated by different mechanisms such as<br />

ischemia–reperfusion, neutrophil and macrophage activation, Fenton chemistry,<br />

endothelial cell xanthine oxidase, free fatty acid and prostaglandin metabolism,<br />

and hypoxia (Figure 5.1) .5,6<br />

Mitochondrial respiration is also one of the main physiological sources for<br />

the generation of ROS. Superoxide is formed when electrons leak from the<br />

electron transport chain. 7 ROS are capable of damaging all biologic macromolecules<br />

including lipids, proteins, polysaccharides, and DNA. When ROS are<br />

overproduced, they are important mediators of cell and tissue injury 8,9 and may<br />

cause cell death by apoptotic and necrotic mechanisms. 10<br />

• Electron transport chain<br />

(mitochondria)<br />

• NADPH oxidase<br />

(neutrophil activation)<br />

• Xanthine oxidase<br />

• Lipid radicals<br />

(fatty acid metabolism)<br />

O 2<br />

NO<br />

.− O2 H 2 O + O 2<br />

GPx, CAT<br />

SOD<br />

ONOO −<br />

H 2 O 2<br />

FIGURE 5.1 Reactive oxygen species produced in tissues.<br />

Fenton reaction<br />

Fe 2+<br />

OH.<br />

Oxidation of DNA,<br />

proteins, and lipids


Oxidative Stress and Antioxidants in the Perinatal Period 73<br />

TABLE 5.1<br />

Major Antioxidants<br />

Enzymes Superoxide dismutase<br />

Catalase<br />

Glutathione peroxidase<br />

Glutathione reductase<br />

Non-enzymatic<br />

antioxidants<br />

Vitamins Vitamin E<br />

Vitamin A<br />

Vitamin C<br />

Coenzyme Q<br />

β-carotene<br />

Reducing agents Glutathione<br />

Cysteine<br />

Thioredoxin<br />

Binding proteins Albumin<br />

Ceruloplasmin<br />

Lactoferrin<br />

Transferrin<br />

Enzyme constituents Zinc<br />

Selenium<br />

Others Uric acid<br />

Bilirubin<br />

Erythropoietin<br />

A more recently appreciated group of molecules known as reactive nitrogen<br />

species are derived from reactions involving another free radical, nitric oxide<br />

(NO). NO can act either as a pro-oxidant or as an antioxidant. 11 O 2· – reacts with<br />

NO to form a cytotoxic oxidant peroxynitrite (ONOO – ) (Figure 5.1).<br />

Oxidative stress can be defined as an imbalance between the amount of ROS<br />

and the intracellular and extracellular antioxidant protection systems. 12 An antioxidant<br />

may be classified as “any substance that can delay or prevent oxidation<br />

of a particular substrate” (Table 5.1) 13 and it may be broadly classified as enzymatic<br />

(superoxide dismutases; SOD, catalase; CAT, glutathione peroxidase; GPx<br />

and glutathione; GSH and their precursors) or non-enzymatic. 14 The antioxidant<br />

defense can be divided into extracellular and intracellular defenses that protect<br />

against ROS-induced cell damage. Transferrin, ceruloplasmin, vitamin C, vitamin<br />

E, uric acid, bilirubin, sulfhydryl groups, and other unidentified antioxidants<br />

contribute to the total antioxidant capacities of extracellular fluids. 15<br />

The extent of oxidative stress has been variably determined by measurement<br />

of a decrease in total antioxidant capacity, through depletion of individual antioxidants<br />

such as vitamin E, vitamin C, or GPx. Otherwise it is defined as the<br />

product (marker) of oxidative stress to lipids, proteins, and DNA.


74 Oxidative Stress and Inflammatory Mechanisms<br />

PREGNANCY: <strong>OXIDATIVE</strong> <strong>STRESS</strong><br />

Pregnancy accompanied by a high metabolic demand and elevated requirements<br />

for tissue oxygen represents a physiological state of oxidative stress. 2 The placenta<br />

has been identified as an important source of lipid peroxidation because of its<br />

enrichment with polyunsaturated fatty acids (PUFAs). 16 Falkay et al. suggested<br />

that the increase in the lipid peroxidation levels is due to increased prostaglandin<br />

synthesis in the placenta. 17 Levels of peroxidation markers such as lipid hydroperoxide<br />

and malondialdehyde (MDA) are higher in pregnant than in non-pregnant<br />

women. 18 Lipid peroxidation is enhanced in the second trimester, tapers off<br />

later in gestation, and decreases after delivery. 19<br />

A certain amount of ROS promotes embryonic development. 20–22 ROS seem<br />

to play a role in signal transduction by modulating transcription factors including<br />

hypoxia-inducible factor (HIF-1) 23 and activated protein-1 (AP-1) 24 that control<br />

the expression of cell growth mediators. ROS may also affect activation and<br />

release of the nuclear factor NFκB — an important regulator of cytokine and<br />

anti-apoptotic gene expression. 23,25,26 However, over-production of ROS may also<br />

induce early embryonic developmental block and retardation (Figure 5.2). 20,23,27<br />

The placenta is also a source of antioxidative enzymes to control placental<br />

lipid peroxidation during uncomplicated pregnancies. All the major antioxidant<br />

systems including SOD, CAT, GPx, glutathione S-transferase, and GSH, and<br />

vitamins C and E are present in the placenta. 28–31 The activities of SOD and CAT<br />

increased as gestation progressed, while the activity of GPx and vitamin E<br />

concentration did not significantly change. 29,32,33 One study found that placental<br />

concentration of lipid peroxidation decreased as gestation advanced, with<br />

increased activities of SOD and CAT in the placenta. 34 Placental antioxidant<br />

defense systems may be sufficient to control the lipid peroxidation in normal<br />

pregnancies. 35,36<br />

Embryonic cell<br />

Transcription factor activation<br />

(HIF-1, AP-1, NF-κ B)<br />

Reactive oxygen species<br />

Oxidative stress<br />

Apoptosis/necrosis<br />

Promote cell development Impair cell development<br />

FIGURE 5.2 Role of reactive oxygen species in fetal development. (Source: Modified<br />

from Dennery, P.A. Antioxid. Redox Signal. 6, 147, 2004.)


Oxidative Stress and Antioxidants in the Perinatal Period 75<br />

Abnormal placentation (inadequate placental invasion of the maternal spiral<br />

arteries) leads to reduced uteroplacental blood flow and placental ischemia. 37 The<br />

ischemia–reperfusion to the placenta leads to abnormal generation of placental<br />

oxidative stress. Excessive placental oxidative stress was assumed to play a role<br />

in the pathogenesis of preeclampsia (PE) 16,38,39 and intrauterine fetal growth<br />

retardation (IUGR). 39–42<br />

PREECLAMPSIA<br />

Preeclampsia (PE) is a pregnancy-specific disorder that complicates 5% of all<br />

pregnancies and 11% of all first pregnancies. 42 Clinically, PE is usually diagnosed<br />

in late pregnancy by increased blood pressure and proteinuria, and the symptoms<br />

of PE typically disappear shortly after delivery of the placenta. The placenta plays<br />

a major role in the pathogenesis of PE, characterized by abnormal placentation<br />

and reduced placental perfusion. 43 Recent evidence suggests that a disturbance<br />

of normal endothelial cell function may be a primary cause in the pathogenesis<br />

of PE. 44,45 PE is associated with severe maternal and fetal mortality 46 and is a<br />

major risk factor for preterm delivery and IUGR. 47<br />

Placental oxidative stress resulting from ischemia–reperfusion injury is<br />

involved in the pathogenesis of PE. A significant increase in placental lipid<br />

peroxidation levels in the placenta of PE has been demonstrated. 48–50 Several<br />

reports confirm that circulating levels of lipid peroxides are significantly elevated<br />

in women with PE compared to normal pregnant women. 18,51–53 However, other<br />

studies showed no evidence of increased circulating lipid peroxidation in PE. 54–56<br />

Lluba et al. speculated that circulating lipid peroxides generated in preeclamptic<br />

women are neutralized by the increasing concentration of antioxidant<br />

enzymes and vitamin E. 55 On the other hand, several important antioxidants are<br />

significantly decreased in pregnant women with PE. The levels of vitamin C,<br />

vitamin E, vitamin A, β-carotene, coenzyme Q10, and GSH are all significantly<br />

lower than in normal pregnancy. 52,53,57,58 Although most studies report decreased<br />

levels of vitamin E in preeclamptic women, some do not. 54,55<br />

Vitamin C and vitamin E have been studied related to prevention of PE. Early<br />

intervention at 16 to 22 weeks of pregnancy (283 women) with vitamin C (1000<br />

mg/day) and E (400 IU/day) supplementation resulted in significant reductions<br />

of PE in the supplement group 59 but a recent report of a randomized trial (100<br />

women) failed to reveal the benefits of PE prevention after the same amounts of<br />

vitamins C and E were supplemented at 14 to 20 weeks of gestation. 60<br />

<strong>OXIDATIVE</strong> <strong>STRESS</strong> AT CHILDBIRTH<br />

Childbirth is an oxidative challenge for newborns. The transition from fetal to<br />

neonatal life at birth implies acute and complex physiologic changes. The fetus<br />

transfers from an intrauterine hypoxic environment with a pO 2 of 20 to 25 mm<br />

Hg to an extrauterine normoxic environment with a pO 2 of 100 mm Hg. 61 This<br />

four- to five-fold increase is believed to induce increased production of ROS.


76 Oxidative Stress and Inflammatory Mechanisms<br />

Neonatal plasma has relatively poor antioxidative defenses. 62 At birth, neonatal<br />

plasma concentrations of vitamins A, E and β-carotene were significantly<br />

lower than maternal plasma levels, while neonatal levels of vitamin C were<br />

significantly higher. 63 Uric acid and vitamin C constitute most of the extracellular<br />

antioxidant capacity, totaling 75%. At 2 weeks of age, these two components<br />

represent only 35% of the extracellular capacity. This change is caused by the<br />

rapid decline of vitamin C levels during the first few days of life and the increasing<br />

concentration of bilirubin which acts as an antioxidant in the first 1 to 2 weeks<br />

post-partum. 64<br />

Term labor is associated with oxidative stress for the neonate. Reported MDA<br />

levels were higher for term infants born by cesarean section than for those born<br />

by spontaneous vaginal delivery. 65,66 In a case controlled study, the serum levels<br />

of lipid peroxidation products were significantly higher (110%) in 20 women<br />

during labor compared to the 20 controls (pregnant women not in labor and<br />

matched for maternal gestational age). 67 On the other hand, women during term<br />

labor showed up-regulations of red blood cell GSH in cord blood. Fetal (difference<br />

in GSH concentration in umbilical vein and artery) and maternal (pre- and postdelivery)<br />

GSH concentrations were significantly lower during labor at term than<br />

in elective cesarean section. 68<br />

PERINATAL ASPHYXIA<br />

Perinatal asphyxia is characterized by transient hypoxia during the ischemic phase<br />

followed by reperfusion. One definition of birth asphyxia is based on the finding<br />

of three of the following four criteria: (1) pH of umbilical arterial cord blood<br />


Oxidative Stress and Antioxidants in the Perinatal Period 77<br />

ATP<br />

Ischemia<br />

Tissue damage<br />

Denosine<br />

Inosine<br />

Hypoxanthine<br />

Reperfusion (restores O 2 )<br />

− Xanthine, O , H2O 2 2<br />

Fe 2+ /Cu 2+<br />

FIGURE 5.3 Suggested mechanism for tissue damage upon reperfusion of hypoxic and<br />

ischemic tissues. (Source: Modified from Halliwell, B. and Gutteridge, J.M., Free Radicals<br />

in Biology and Medicine, Oxford University Press, New York, 1999.)<br />

approached term. 40 ROS cause endothelial cell damage and abnormalities in Nmethyl-D-aspartate<br />

receptors, synaptosome structures, and astrocyte functions,<br />

thus contributing to the development of brain injury after a hypoxic–ischemic<br />

episode. 74–76<br />

USE OF 100% OXYGEN IN RESUSCITATION<br />

Xanthine dehydrogenase<br />

Xanthine oxidase<br />

The optimal concentration of oxygen for neonatal resuscitation is uncertain.<br />

Traditionally, neonatal resuscitation has been performed with 100% oxygen.<br />

Many textbooks and the advisory statement of the International Liaison Committee<br />

on Resuscitation 77 recommend that resuscitation of newborn infants should<br />

be performed with 100% oxygen. Resuscitation of depressed newborns with<br />

presumed hypoxia and/or asphyxia in the delivery room is currently the only<br />

remaining clinical indication for the use of unregulated 100% oxygen in infants.<br />

Exposure to high levels of oxygen during reoxygenation may promote the<br />

formation of excessive levels of ROS in tissue, causing tissue injury. Hyperoxemia<br />

has been associated with numerous negative side effects including delayed initiation<br />

of spontaneous respiration, increased oxygen consumption, and irregularities<br />

in cerebral circulation. 78 Therefore, several groups have investigated the safety<br />

and efficacy of using room air for neonatal resuscitation.<br />

Evidence from animal studies proves that room air is as effective as 100%<br />

oxygen in resuscitation. No significant differences were noted between two<br />

groups (hypoxemic newborn pigs resuscitated with 21% O 2 or 100% O 2 for 20<br />

to 25 min followed by 21% O 2) in arterial blood pressure, base deficit, plasma<br />

OH −


78 Oxidative Stress and Inflammatory Mechanisms<br />

GSH/GSSG<br />

80<br />

60<br />

40<br />

20<br />

Nonasphyxiated neonates<br />

RAR<br />

OxR<br />

0<br />

0 3 28<br />

Days of postnatal life<br />

FIGURE 5.4 Reduced gluthatione (GSH) and oxidized gluthathione (GSSG) ratio in<br />

asphyxiated neonates resuscitated with 100% oxygen (OxR) or room air (RAR). ** p<br />


Oxidative Stress and Antioxidants in the Perinatal Period 79<br />

antioxidant synthesis and activities) are low in preterm infants compared to term<br />

infants. 88,89 Induction of antioxidative enzymes following oxidative stress was<br />

found in term but not in preterm infants. 90 These data strongly suggest increased<br />

susceptibilities of premature infants to oxidative stress. 62 Clinical conditions in<br />

which oxidative stress may be particularly relevant in premature infants are<br />

discussed below.<br />

BRONCHOPULMONARY DYSPLASIA<br />

Ventilated premature infants are at risk of developing bronchopulmonary dysplasia<br />

(BPD), a chronic lung disease of prematurity. When BPD was first described,<br />

exposure to a high oxygen level was identified as a risk factor for its<br />

development 91,92 and it was soon related to ROS. 93 It is widely assumed that BPD<br />

is due to ROS-related injury to the immature lung. 94 Two studies found a close<br />

relationship between high oxygen exposure and the development of BPD. 95,96 Van<br />

Marter et al. found that inspired oxygen concentrations were higher in infants<br />

who developed BPD compared with those who did not. A fractional inspired<br />

oxygen level of 1.0 in the first day of life almost doubled the risk of BPD relative<br />

to an FiO 2. 96 Oxidative stress detected by the augmentation of lipid peroxidation<br />

products in the very first days of life is associated with the subsequent development<br />

of BPD. 97,98<br />

Antioxidants present in the alveolar epithelial lining fluid are well positioned<br />

to protect against tissue damage caused by ROS generated by these mechanisms. 99<br />

Infants who develop BPD have decreased concentrations of glutathione in bronchoalveolar<br />

fluid compared to those who did not require supplemental oxygen at<br />

36 weeks post-conceptional age on the first day of life. 86 Robbins et al. reported<br />

that the administration of recombinant human SOD (rhSOD) in newborn piglets<br />

mitigated the lung inflammatory changes (analyzed by BAL neutrophil chemotactic<br />

activity and cell count), MDA concentration in lung tissue, and acute lung<br />

injury induced by 100 ppm of NO and 90% O 2. 100 In a multicenter controlled<br />

study of 26 preterm infants, significant reductions in radiological evidence of<br />

BPD were noted in infants treated with rhSOD compared to infants treated with<br />

placebo. 101 However, in a later and larger study of 302 infants, no differences in<br />

the development of BPD in infants treated with rhSOD and those treated with<br />

placebo were noted. 102<br />

NO is a free radical that may be oxidized or reduced, depending on its<br />

concentration and the presence of other oxidants such as oxygen. Hyperoxic<br />

exposure of rat pups up-regulated both inducible and endothelial NO synthase<br />

and therefore increased the concentrations of NO and subsequently peroxynitrite<br />

as well. 103 NO is highly toxic for preterm infants. However, a clinical trial with<br />

inhaled NO did not report higher occurrences of BPD in NO-treated premature<br />

infants compared with controls. 104 Hamon et al. recently reported that low doses<br />

(5 ppm) of inhaled NO administered soon after birth in hypoxemic premature<br />

infants were associated with significant decreases in their oxidative stress after<br />

24 hr as assessed by plasma MDA. 105


80 Oxidative Stress and Inflammatory Mechanisms<br />

NEONATAL NECROTIZING ENTEROCOLITIS<br />

Neonatal necrotizing enterocolitis (NEC) is the most common life-threatening<br />

disease of the gastrointestinal tract in the neonatal period; it primarily affects<br />

premature infants. 106 NEC is characterized by various degrees of mucosal or<br />

transmural necrosis of the intestine. Its causes remain unclear but are most likely<br />

multifactorial. Intestinal ischemia and colonization of the intestinal lumen by<br />

several viral and bacterial organisms were hypothesized to cause NEC. 107 Immunologic<br />

immaturity of the preterm gut 108 and neonatal hypoxia were also included<br />

as causative factors. Upon reperfusion and reoxygenation of the damaged intestine,<br />

a flood of ROS is generated by the xanthine oxidase system. This burst of<br />

ROS can cause severe tissue damage and may be the final pathway for tissue<br />

injury of the gut mucosa noted in NEC. 71<br />

ROS can alter inflammation in the gut and lead to activation of platelet<br />

activating factor (PAF), an endogenous phospholipid mediator 109 that results in<br />

increased coagulation and subsequent microthrombus formation in small vessels.<br />

A model of acute inflammatory intestinal injury induced by PAF was significantly<br />

attenuated by administration of allopurinol (a xanthine inhibitor). 110<br />

PERIVENTRICULAR LEUKOMALACIA<br />

Periventricular leukomalacia (PVL) in premature infants is a distinctive lesion of<br />

cerebral white matter (Figure 5.5) associated with severe adverse neurologic<br />

outcomes. The pathogenesis of cerebral white matter injury in premature infants<br />

FIGURE 5.5 Coronal sections of brain with periventricular leukomalacia.


Oxidative Stress and Antioxidants in the Perinatal Period 81<br />

is not entirely clear, although ischemia-reperfusion and infection-inflammation<br />

appear important. 70 The brain is at special risk due to its high concentration of<br />

PUFAs and deficiency of SOD and GPx. 111 Immature oligodendrocytes are more<br />

prone to oxidative stress than mature ones. 112 These immature forms, the so-called<br />

preoligodendrocytes, have been shown recently to account for 90% of the total<br />

population of oligodendrocytes in cerebral white matter of infants under the<br />

gestational age of 31 weeks 113 who represent a high risk group for the development<br />

of PVL. 114<br />

Houdou et al. also reported that CAT-positive glia did not appear in the deep<br />

white matter before 31 to 32 weeks of gestation. 115 These very premature infants<br />

are most at risk for white matter injury. 116 In two model systems of free radical<br />

accumulation, the early differentiating oligodendrocyte was shown to be exquisitely<br />

vulnerable to free radical attack. 113,117<br />

Direct support for an association between products of oxidation and PVL in<br />

premature infants is limited. However, in a recent study, premature infants with<br />

subsequent evidence of PVL on magnetic resonance imaging at term had higher<br />

levels of cerebrospinal fluid protein carbonyls (markers of oxidized protein) than<br />

healthy premature or term infants. 116<br />

RETINOPATHY OF PREMATURITY<br />

Retinopathy of prematurity (ROP) is a vasoproliferative retinal disorder of premature<br />

infants. Its basic pathogenesis is not fully understood but exposure to the<br />

extrauterine environment including necessarily high inspired oxygen concentrations<br />

produces cellular damage mediated by ROS. Hyperoxygenation favors peroxidation<br />

of vasoactive isoprostanes, resulting in vasoconstriction and vascular<br />

cytotoxicity leading to ischemia, which predisposes to the development of vasoproliferative<br />

retinopathy. 118 Severe ROP can lead to lifelong visual impairment<br />

or blindness. 119<br />

Early animal models of ROP first suggested that oxygen was involved in the<br />

normal developing retinal vasculature, probably via a putative angiogenic factor 120<br />

that was subsequently identified as vascular endothelial growth factor (VEGF). 121<br />

VEGF plays an important role in endothelial cell proliferation, migration, and<br />

blood vessel formation 122,123 and in the development of ROP. 124<br />

Several studies have shown that greater variability of transcutaneous oxygen<br />

during the first 2 weeks of life is associated with the development of severe<br />

ROP. 125,126 The so-called STOP-ROP (supplemental therapeutic oxygen for prethreshold<br />

retinopathy of prematurity) trial enrolled about 600 premature infants<br />

with confirmed threshold ROP in at least one eye. The risks of adverse pulmonary<br />

events including BPD increased with the use of supplemental oxygen at pulse<br />

oximetry saturation of 96 to 99% compared with pulse oximetry targeting 89 to<br />

94%. However, this therapy had no significant effect on the progression of ROP. 127<br />

Supplementation of antioxidants has been reported to protect against ROP. A<br />

meta-analysis evaluated data from six randomized clinical trials of vitamin E<br />

prophylaxis (15 to 100 mg/kg/day from day 1 until discharge) that included a


82 Oxidative Stress and Inflammatory Mechanisms<br />

total of 704 premature infants in the vitamin E prophylaxis groups and 714 control<br />

infants. The overall incidence of any stage ROP was similar between the groups,<br />

but incidence of severe (stage +3) ROP was lower in the vitamin E-treated group<br />

(2.4% in the vitamin E group versus 5.3% of controls). The authors concluded<br />

that the role of the vitamin E antioxidant in reducing severe ROP must be reevaluated.<br />

128<br />

HUMAN MILK <strong>AND</strong> ANTI<strong>OXIDATIVE</strong> PROTECTION<br />

Human milk (HM) is considered the ideal food for healthy infants 129 and is known<br />

to contain various bioactive substances, some of which are reported to be antioxidants.<br />

130 HM contains many antioxidants such as enzymes (CAT, GPx, SOD),<br />

vitamins (A, C, E), binding proteins such as lactoferrin, and constituents of<br />

antioxidative enzymes (Cu, Zn). 131–133 In contrast, antioxidative enzymes are<br />

absent from infant formula. 134 Most infant formula has higher amounts of vitamins<br />

added than are present in HM to make up for the reduced bioavailability. Thus,<br />

the overall antioxidant capacity of HM versus infant formula is difficult to assess,<br />

although assessment would probably favor HM. 14<br />

We previously reported in vitro results showing that HM alleviated H 2O 2induced<br />

oxidative damage in intestinal epithelial cell lines, whereas bovine milk<br />

or infant formula did not. 135 Confluent intestinal epithelial (IEC-6) cells were<br />

preincubated with defatted HM, bovine milk, or three artificial milks for 24 hr<br />

followed by H 2O 2 challenge. HM-treated cells showed the highest survival rates<br />

(50%) compared with bovine milk-treated (6%) or infant formula-treated (13 to<br />

16%) cells (Figure 5.6). 135 Buescher and Mcllherhan 131 reported that human<br />

colostrum manifested antioxidant properties, proving capable of spontaneous<br />

reduction of cytochrome C, depletion of polymorphonuclear leukocyte-produced<br />

H 2O 2, and protection of epithelial cells from polymorphonuclear leukocyte-mediated<br />

detachment.<br />

Several clinical studies demonstrated antioxidative properties of HM. HMfed<br />

infants had higher plasma trapping ability (a measure of resistance to oxidative<br />

stress in vitro) than did control infants fed formula. 136 We compared the oxidative<br />

stress levels in 41 healthy 1-month-old infants and 29 premature infants by<br />

measuring urinary 8-hydroxy-2-deoxyguanosine (8-OHdG — a marker of oxidative<br />

DNA damage). In the 1-month-old group, urinary 8-OHdG excretions of the<br />

breast-fed infants were significantly lower than those of the artificial formula-fed<br />

infants (Figure 5.7). 137 In the premature group, urinary 8-OHdG excretions of the<br />

breast-fed infants at 14 and 28 days of age were significantly lower than those<br />

of the formula-fed infants (Figure 5.8). 138 These data indicate that HM provides<br />

more antioxidant properties than infant formula during early infancy. This may<br />

be due to the presence of antioxidants in HM that may exhibit antioxidant effects<br />

in the gut and may pass through the relatively porous neonatal intestine early in<br />

infancy. 131,132 In fact, feeding with HM has been associated with a lower incidence<br />

of a variety of illnesses including NEC, 139 respiratory disease, 140 and ROP 141 in<br />

premature infants.


Oxidative Stress and Antioxidants in the Perinatal Period 83<br />

Cell viability (% of control)<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

No pretreatment<br />

Human<br />

colostrum<br />

Infant formulas<br />

(a) (b) (c)<br />

Cow’s<br />

milk<br />

FIGURE 5.6 Antioxidative properties of human colostrum, bovine milk, and artificial<br />

formulas and H 2O 2-induced oxidative damage in IEC-6 cells. Survival cell rates are<br />

expressed as percentages of control (non-H 2O 2-challenged) cells. Values are means + SD.<br />

* p formula<br />

group received 50 to 90% of their intake as breast milk. Formula>breast group received<br />

50 to 90% of their intake as formula. Formula-fed group received 90% of their intake as<br />

formula. Values = means + SD. * p


84 Oxidative Stress and Inflammatory Mechanisms<br />

Urinary 8-OHdG<br />

excretion (ng/mg . Cr)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

∗<br />

Day 14<br />

FIGURE 5.8 Change in urinary 8-hydroxydeoxyguanosine (8-OHdG) excretion at 14 and<br />

28 days of age in breast-fed and formula-fed very low birthweight infants. Values = means<br />

+ SD. *p


Oxidative Stress and Antioxidants in the Perinatal Period 85<br />

way anticipates future policies in this area. BK is the recipient of a Freedom to<br />

Discover Award of the Bristol Myers Squibb Foundation, New York, NY, U.S.A.<br />

REFERENCES<br />

1. Halliwell, B. and Gutteridge, J.M. Free Radicals in Biology and Medicine, Oxford<br />

University Press, New York, 1999.<br />

2. Spatling, L., Fallenstein, F., Huch, A., Huch, R., and Rooth, G. The variability of<br />

cardiopulmonary adaptation to pregnancy at rest and during exercise. Br. J. Obstet.<br />

Gynaecol. 99, Suppl. 8, 1, 1992.<br />

3. Granot, E. and Kohen, R. Oxidative stress in childhood — in health and disease<br />

states. Clin. Nutr. 23, 3, 2004.<br />

4. Saugstad, O.D. Hypoxanthine as an indicator of hypoxia: its role in health and<br />

disease through free radical production. Pediatr. Res. 23, 143, 1988.<br />

5. McCord, J.M. Oxygen-derived free radicals in postischemic tissue injury. New<br />

Engl. J. Med. 312, 159, 1985.<br />

6. Mishra, O.P. and Delivoria-Papadopoulos, M. Cellular mechanisms of hypoxic<br />

injury in the developing brain. Brain Res. Bull. 48, 233, 1999.<br />

7. Halliwell, B., Gutteridge, J.M., and Cross, C.D. Free radicals, antioxidants, and<br />

human disease: where are we now? J. Lab. Clin. Med. 119, 598, 1992.<br />

8. Fridovich, I. Oxygen toxicity: a radical explanation. J. Exp. Biol. 201, 1203, 1998.<br />

9. Halliwell, B. Free radicals, antioxidants, and human disease: curiosity, cause, or<br />

consequence? Lancet 344, 721, 1994.<br />

10. Wiseman, H. and Halliwell, B. Damage to DNA by reactive oxygen and nitrogen<br />

species: role in inflammatory disease and progression to cancer. Biochem. J. 313,<br />

Pt. 1, 17, 1996.<br />

11. Patel, R.P. et al. Biological aspects of reactive nitrogen species. Biochim. Biophys.<br />

Acta 1411, 385, 1999.<br />

12. Sies, H. Oxidative stress: oxidants and antioxidants. Exp. Physiol. 82, 291, 1997.<br />

13. Halliwell, B. The antioxidant paradox. Lancet 355, 1179, 2000.<br />

14. Rassin, D.K. and Smith, K.E. Nutritional approaches to improve cognitive development<br />

during infancy: antioxidant compounds. Acta Paediatr. Suppl. 92, 34,<br />

2003.<br />

15. Jankov, R.P., Negus, A., and Tanswell, A.K. Antioxidants as therapy in the newborn:<br />

some words of caution. Pediatr. Res. 50, 681, 2001.<br />

16. Mutlu-Turkoglu, U. et al. Imbalance between lipid peroxidation and antioxidant<br />

status in preeclampsia. Gynecol. Obstet. Invest. 46, 37, 1998.<br />

17. Falkay, G., Herczeg, J., and Sas, M. Microsomal lipid peroxidation in human<br />

pregnant uterus and placenta. Biochem. Biophys. Res. Commun. 79, 843, 1977.<br />

18. Morris, J.M. et al. Circulating markers of oxidative stress are raised in normal<br />

pregnancy and pre-eclampsia. Br. J. Obstet. Gynaecol. 105, 1195, 1998.<br />

19. Little, R.E. and Gladen, B.C. Levels of lipid peroxides in uncomplicated pregnancy:<br />

a review of the literature. Reprod. Toxicol. 13, 347, 1999.<br />

20. Dennery, P.A. Role of redox in fetal development and neonatal diseases. Antioxid.<br />

Redox. Signal. 6, 147, 2004.<br />

21. Fisher, J.C. et al. Oxidation–reduction (redox) controls fetal hypoplastic lung<br />

growth. J. Surg. Res. 106, 287, 2002.


86 Oxidative Stress and Inflammatory Mechanisms<br />

22. Harvey, A.J., Kind, K.L., and Thompson, J.G.. Redox regulation of early embryo<br />

development. Reproduction 123, 479, 2002.<br />

23. Haddad, J.J., Olver, R.E., and Land, S.C. Antioxidant/pro-oxidant equilibrium<br />

regulates HIF-1 alpha and NF-kappa B redox sensitivity: evidence for inhibition<br />

by glutathione oxidation in alveolar epithelial cells. J. Biol. Chem. 275, 21130,<br />

2000.<br />

24. Gomez, D.A. et al. Antioxidants and AP-1 activation: a brief overview. Immunobiology<br />

198, 273, 1997.<br />

25. Burdon, R.H. Superoxide and hydrogen peroxide in relation to mammalian cell<br />

proliferation. Free Radical Biol. Med. 18, 775, 1995.<br />

26. Mattson, M.P. et al. Roles of nuclear factor kappa-B in neuronal survival and<br />

plasticity. J. Neurochem. 74, 443, 2000.<br />

27. Bedaiwy, M.A. et al. Differential growth of human embryos in vitro: role of<br />

reactive oxygen species. Fertil. Steril. 82, 593, 2004.<br />

28. Guthenberg, C. and Mannervik, B. Glutathione S-transferase (transferase pi) from<br />

human placenta is identical or closely related to glutathione S-transferase (transferase<br />

rho) from erythrocytes. Biochim. Biophys. Acta 661, 255, 1981.<br />

29. Myatt, L. and Cui, X. Oxidative stress in the placenta. Histochem. Cell. Biol. 122,<br />

369, 2004.<br />

30. Van Hien, P., Kovacs, K., and Matkovics, B. Properties of enzymes. I. Study of<br />

superoxide dismutase activity change in human placenta of different ages. Enzyme<br />

18, 341, 1974.<br />

31. Wang, Y. and Walsh, S.W. Antioxidant activities and mRNA expression of superoxide<br />

dismutase, catalase, and glutathione peroxidase in normal and preeclamptic<br />

placentas. J. Soc. Gynecol. Investig. 3, 179, 1996.<br />

32. Watson, A.L. et al. Variations in expression of copper/zinc superoxide dismutase<br />

in villous trophoblast of the human placenta with gestational age. Placenta 18,<br />

295, 1997.<br />

33. Spinnato, J.A. and Livingston, J.C. Prevention of preeclampsia with antioxidants:<br />

evidence from randomized trials. Clin. Obstet. Gynecol. 48, 416, 2005.<br />

34. Qanungo, S. and Mukherjea, M. Ontogenic profile of some antioxidants and lipid<br />

peroxidation in human placental and fetal tissues. Mol. Cell Biochem. 215, 11,<br />

2000.<br />

35. Gitto, E. et al. Causes of oxidative stress in the pre- and perinatal period. Biol.<br />

Neonate 81, 146, 2002.<br />

36. Mueller, A. et al. Placental defence is considered sufficient to control lipid peroxidation<br />

in pregnancy. Med. Hypotheses 64, 553, 2005.<br />

37. van Beck, E. and Peters, L.L. Pathogenesis of preeclampsia: a comprehensive<br />

model. Obstet. Gynecol. Surv. 53, 233, 1998.<br />

38. Gupta, S., Agarwal, A., and Sharma, R.K. The role of placental oxidative stress<br />

and lipid peroxidation in preeclampsia. Obstet. Gynecol. Surv. 60, 807, 2005.<br />

39. Takagi, Y. et al. Levels of oxidative stress and redox-related molecules in the<br />

placenta in preeclampsia and fetal growth restriction. Virchows Arch. 444, 49,<br />

2004.<br />

40. Mishra, O.P. and Delivoria-Papadopoulos, M. Modification of modulatory sites<br />

of NMDA receptor in the fetal guinea pig brain during development. Neurochem.<br />

Res. 17, 1223, 1992.<br />

41. Scholl, T.O. and Stein, T.P. Oxidant damage to DNA and pregnancy outcome. J.<br />

Matern. Fetal Med. 10, 182, 2001.


Oxidative Stress and Antioxidants in the Perinatal Period 87<br />

42. Agarwal, A., Gupta, S., and Sharma, R.K. Role of oxidative stress in female<br />

reproduction. Reprod. Biol. Endocrinol. 3, 28, 2005.<br />

43. Roberts, J.M. Preeclampsia: what we know and what we do not know. Semin.<br />

Perinatol. 24, 24, 2000.<br />

44. Hubel, C.A. Oxidative stress in the pathogenesis of preeclampsia. Proc. Soc. Exp.<br />

Biol. Med. 222, 222, 1999.<br />

45. VanWijk, M.J. et al. Vascular function in preeclampsia. Cardiovasc. Res. 47, 38,<br />

2000.<br />

46. Villar, J. et al. Methodological and technical issues related to the diagnosis,<br />

screening, prevention, and treatment of pre-eclampsia and eclampsia. Int. J.<br />

Gynaecol. Obstet. 85, Suppl. 1, S28, 2004.<br />

47. Lim, K.H. and Friedman, S.A. Hypertension in pregnancy. Curr. Opin. Obstet.<br />

Gynecol. 5, 40, 1993.<br />

48. Atamer, Y. et al. Lipid peroxidation, antioxidant defense, status of trace metals<br />

and leptin levels in preeclampsia. Eur. J. Obstet. Gynecol. Reprod. Biol. 119, 60,<br />

2005.<br />

49. Vanderlelie, J. et al. Increased biological oxidation and reduced anti-oxidant<br />

enzyme activity in pre-eclamptic placentae. Placenta 26, 53, 2005.<br />

50. Walsh, S.W. et al. Placental isoprostane is significantly increased in preeclampsia.<br />

FASEB J. 14, 1289, 2000.<br />

51. Maseki, M. et al. Lipid peroxide levels and lipids content of serum lipoprotein<br />

fractions of pregnant subjects with or without pre-eclampsia. Clin. Chim. Acta<br />

115, 155, 1981.<br />

52. Wang, Y.P. et al. Imbalance between thromboxane and prostacyclin in preeclampsia<br />

is associated with an imbalance between lipid peroxides and vitamin E in<br />

maternal blood. Am. J. Obstet. Gynecol. 165, 1695, 1991.<br />

53. Yanik, F.F. et al. Pre-eclampsia associated with increased lipid peroxidation and<br />

decreased serum vitamin E levels. Int. J. Gynaecol. Obstet. 64, 27, 1999.<br />

54. Bowen, R.S. et al. Oxidative stress in pre-eclampsia. Acta Obstet. Gynecol. Scand.<br />

80, 719, 2001.<br />

55. Llurba, E. et al. A comprehensive study of oxidative stress and antioxidant status<br />

in preeclampsia and normal pregnancy. Free Radic. Biol. Med. 37, 557, 2004.<br />

56. Regan, C.L. et al. No evidence for lipid peroxidation in severe preeclampsia. Am.<br />

J. Obstet. Gynecol. 185, 572, 2001.<br />

57. Mikhail, M.S. et al. Preeclampsia and antioxidant nutrients: decreased plasma<br />

levels of reduced ascorbic acid, alpha-tocopherol, and beta-carotene in women<br />

with preeclampsia. Am. J. Obstet. Gynecol. 171, 150, 1994.<br />

58. Palan, P.R. et al. Lipid-soluble antioxidants and pregnancy: maternal serum levels<br />

of coenzyme Q10, alpha-tocopherol and gamma-tocopherol in preeclampsia and<br />

normal pregnancy. Gynecol. Obstet. Invest. 58, 8, 2004.<br />

59. Chappell, L.C. et al. Effect of antioxidants on the occurrence of pre-eclampsia in<br />

women at increased risk: a randomised trial. Lancet 354, 810, 1999.<br />

60. Beazley, D. et al. Vitamin C and E supplementation in women at high risk for<br />

preeclampsia: a double-blind, placebo-controlled trial. Am. J. Obstet. Gynecol.<br />

192, 520, 2005.<br />

61. Muller, D.P. Free radical problems of the newborn. Proc. Nutr. Soc. 46, 69, 1987.<br />

62. Saugstad, O.D. Oxidative stress in the newborn: a 30-year perspective. Biol.<br />

Neonate 88, 228, 2005.


88 Oxidative Stress and Inflammatory Mechanisms<br />

63. Scaife, A.R. et al. Maternal intake of antioxidant vitamins in pregnancy in relation<br />

to maternal and fetal plasma levels at delivery. Br. J. Nutr. 95, 771, 2006.<br />

64. Berger, H.M. et al. Extracellular defence against oxidative stress in the newborn.<br />

Semin. Neonatol. 3, 183, 1998.<br />

65. Rogers, M.S. et al. Lipid peroxidation in cord blood at birth: the effect of labour.<br />

Br. J. Obstet. Gynaecol. 105, 739, 1998.<br />

66. Yaacobi, N., Ohel, G., and Hochman, A. Reactive oxygen species in the process<br />

of labor. Arch. Gynecol. Obstet. 263, 23, 1999.<br />

67. Fainaru, O. et al. Active labour is associated with increased oxidisibility of serum<br />

lipids ex vivo. Br. J. Obstet. Gynaecol. 109, 938, 2002.<br />

68. Buhimschi, I.A. et al. Beneficial impact of term labor: nonenzymatic antioxidant<br />

reserve in the human fetus. Am. J. Obstet. Gynecol. 189, 181, 2003.<br />

69. Committee on Fetus and Newborn, American Academy of Pediatrics, and Committee<br />

on Obstetric Practice, American College of Obstetricians and Gynecologists.<br />

Use and abuse of the Apgar score. Pediatrics 98, 141, 1996.<br />

70. Volpe, J.J. Neurology of the Newborn, W.B. Saunders, Philadelphia, 2001, p. 314.<br />

71. Jassem, W. and Roake, J. The molecular and cellular basis of reperfusion injury<br />

following organ transplantation. Transplantation Rev. 12, 14, 1998.<br />

72. Buonocore, G., Perrone, S., and Bracci, R. Free radicals and brain damage in the<br />

newborn. Biol. Neonate 79, 180, 2001.<br />

73. Mishra, O.P. and Delivoria-Papadopoulos, M. Lipid peroxidation in developing<br />

fetal guinea pig brain during normoxia and hypoxia. Brain Res. Dev. Brain Res.<br />

45, 129, 1989.<br />

74. Berger, R. and Garnier, Y. Perinatal brain injury. J. Perinat. Med. 28, 261, 2000.<br />

75. Bracci, R., Perrone, S., and Buonocore, G. Red blood cell involvement in fetal/neonatal<br />

hypoxia. Biol. Neonate 79, 210, 2001.<br />

76. Fellman, V. and Raivio, K.O. Reperfusion injury as the mechanism of brain<br />

damage after perinatal asphyxia. Pediatr. Res. 41, 599, 1997.<br />

77. Kattwinkel, J. et al. Resuscitation of the newly born infant: an advisory statement<br />

from the Pediatric Working Group of the International Liaison Committee on<br />

Resuscitation. Resuscitation 40, 71, 1999.<br />

78. Saugstad, O.D. Resuscitation with room-air or oxygen supplementation. Clin.<br />

Perinatol. 25, 741, 1998.<br />

79. Rootwelt, T. et al. Hypoxemia and reoxygenation with 21% or 100% oxygen in<br />

newborn pigs: changes in blood pressure, base deficit, and hypoxanthine and brain<br />

morphology. Pediatr. Res. 32, 107, 1992.<br />

80. Rootwelt, T. et al. Cerebral blood flow and evoked potentials during reoxygenation<br />

with 21 or 100% O2 in newborn pigs. J. Appl. Physiol. 75, 2054, 1993.<br />

81. Vento, M. et al. Resuscitation with room air instead of 100% oxygen prevents<br />

oxidative stress in moderately asphyxiated term neonates. Pediatrics 107, 642,<br />

2001.<br />

82. Saugstad, O.D., Rootwelt, T., and Aalen, O. Resuscitation of asphyxiated newborn<br />

infants with room air or oxygen: an international controlled trial. Pediatrics 102,<br />

E1, 1998.<br />

83. Saugstad, O.D. et al. Resuscitation of newborn infants with 21% or 100% oxygen:<br />

follow-up at 18 to 24 months. Pediatrics 112, 296, 2003.<br />

84. Saugstad, O.D. Bronchopulmonary dysplasia, oxidative stress and antioxidants.<br />

Semin. Neonatol. 8, 39, 2003.


Oxidative Stress and Antioxidants in the Perinatal Period 89<br />

85. Phylactos, A.C. et al. Erythrocyte cupric/zinc superoxide dismutase exhibits<br />

reduced activity in preterm and low-birthweight infants at birth. Acta Paediatr.<br />

84, 1421, 1995.<br />

86. Grigg, J., Barber, A., and Silverman, M. Bronchoalveolar lavage fluid glutathione<br />

in intubated premature infants. Arch. Dis. Child. 69, 49, 1993.<br />

87. Jain, A. et al. Glutathione metabolism in newborns: evidence for glutathione<br />

deficiency in plasma, bronchoalveolar lavage fluid, and lymphocytes in prematures.<br />

Pediatr. Pulmonol. 20, 160, 1995.<br />

88. Baydas, G. et al. Antioxidant vitamin levels in term and preterm infants and their<br />

relation to maternal vitamin status. Arch. Med. Res. 33, 276, 2002.<br />

89. Galinier, A. et al. Reference range for micronutrients and nutritional marker<br />

proteins in cord blood of neonates appropriated for gestational ages. Early Hum.<br />

Dev. 81, 583, 2005.<br />

90. Frank, L. and Sosenko, I.R. Failure of premature rabbits to increase antioxidant<br />

enzymes during hyperoxic exposure: increased susceptibility to pulmonary oxygen<br />

toxicity compared with term rabbits. Pediatr. Res. 29, 292, 1991.<br />

91. Bancalari, E. and Gerhardt, T. Bronchopulmonary dysplasia. Pediatr. Clin. North<br />

Am. 33, 1, 1986.<br />

92. Northway, W.H., Jr. and Rosan, R.C. Radiographic features of pulmonary oxygen<br />

toxicity in the newborn: bronchopulmonary dysplasia. Radiology 91, 49, 1968.<br />

93. Bonta, V.W., Gawron, E.R., and Warshaw, J.B. Neonatal red cell superoxide<br />

dismutase enzyme levels: possible role as a cellular defense mechanism against<br />

pulmonary oxygen toxicity. Pediatr. Res. 11, 754, 1977.<br />

94. Abman, S.H. and Groothius, J.R. Pathophysiology and treatment of bronchopulmonary<br />

dysplasia: current issues. Pediatr. Clin. North Am. 41, 277, 1994.<br />

95. Gaynon, M.W. and Stevenson, D.K. What can we learn from STOP-ROP and<br />

earlier studies? Pediatrics 105, 420, 2000.<br />

96. Van Marter, L.J. et al. Do clinical markers of barotrauma and oxygen toxicity<br />

explain interhospital variation in rates of chronic lung disease? Pediatrics 105,<br />

1194, 2000.<br />

97. Ogihara, T. et al. Raised concentrations of aldehyde lipid peroxidation products<br />

in premature infants with chronic lung disease. Arch. Dis. Child. Fetal Neonatal<br />

Ed. 80, F21, 1999.<br />

98. Saugstad, O.D. Chronic lung disease: the role of oxidative stress. Biol. Neonate<br />

74, Suppl. 1, 21, 1998.<br />

99. van Klaveren, R.J., Demedts, M., and Nemery, B. Cellular glutathione turnover<br />

in vitro, with emphasis on type II pneumocytes. Eur. Resp. J. 10, 1392, 1997.<br />

100. Robbins, C.G. et al. Recombinant human superoxide dismutase reduces lung<br />

injury caused by inhaled nitric oxide and hyperoxia. Am. J. Physiol. 272, L903,<br />

1997.<br />

101. Rosenfeld, W.N. et al. Safety and pharmacokinetics of recombinant human superoxide<br />

dismutase administered intratracheally to premature neonates with respiratory<br />

distress syndrome. Pediatrics 97, 811, 1996.<br />

102. Davis, J.M. et al. Pulmonary outcome at 1 year corrected age in premature infants<br />

treated at birth with recombinant human CuZn superoxide dismutase. Pediatrics<br />

111, 469, 2003.<br />

103. Potter, C.F. et al. Effects of hyperoxia on nitric oxide synthase expression, nitric<br />

oxide activity, and lung injury in rat pups. Pediatr. Res. 45, 8, 1999.


90 Oxidative Stress and Inflammatory Mechanisms<br />

104. Lipkin, P.H. et al. Neurodevelopmental and medical outcomes of persistent pulmonary<br />

hypertension in term newborns treated with nitric oxide. J. Pediatr. 140,<br />

306, 2002.<br />

105. Hamon, I. et al. Early inhaled nitric oxide improves oxidative balance in very<br />

preterm infants. Pediatr. Res. 57, 637, 2005.<br />

106. Behrman, R.E. et al. Nelson’s Textbook of Pediatrics, W.B. Saunders, Philadelphia,<br />

2003.<br />

107. Kosloske, A.M. Pathogenesis and prevention of necrotizing enterocolitis: a hypothesis<br />

based on personal observation and a review of the literature. Pediatrics 74,<br />

1086, 1984.<br />

108. Kosloske, A.M. Epidemiology of necrotizing enterocolitis. Acta Paediatr. Suppl.<br />

396, 2, 1994.<br />

109. Bhatia, A.M., Feddersen, R.M., and Musemeche, C.A. Role of luminal nutrients<br />

in intestinal injury from mesenteric reperfusion and platelet-activating factor in<br />

the developing rat. J. Surg. Res. 63, 152, 1996.<br />

110. Cueva, J.P. and Hsueh, W. Role of oxygen derived free radicals in platelet activating<br />

factor induced bowel necrosis. Gut 29, 1207, 1988.<br />

111. Inder, T.E. et al. Lipid peroxidation as a measure of oxygen free radical damage<br />

in the very low birthweight infant. Arch. Dis. Child. Fetal Neonatal Ed. 70, F107,<br />

1994.<br />

112. Baud, O. et al. Glutathione peroxidase-catalase cooperativity is required for resistance<br />

to hydrogen peroxide by mature rat oligodendrocytes. J. Neurosci. 24, 1531,<br />

2004.<br />

113. Back, S.A. et al. Maturation-dependent vulnerability of oligodendrocytes to oxidative<br />

stress-induced death caused by glutathione depletion. J. Neurosci. 18, 6241,<br />

1998.<br />

114. Back, S.A. et al. Late oligodendrocyte progenitors coincide with the developmental<br />

window of vulnerability for human perinatal white matter injury. J. Neurosci.<br />

21, 1302, 2001.<br />

115. Houdou, S. et al. Developmental immunohistochemistry of catalase in the human<br />

brain. Brain Res. 556, 267, 1991.<br />

116. Inder, T. et al. Elevated free radical products in cerebrospinal fluid of VLBW<br />

infants with cerebral white matter injury. Pediatr. Res. 52, 213, 2002.<br />

117. Oka, A. et al. Vulnerability of oligodendroglia to glutamate: pharmacology, mechanisms,<br />

and prevention. J. Neurosci. 13, 1441, 1993.<br />

118. Hardy, P. et al. Oxidants, nitric oxide and prostanoids in the developing ocular<br />

vasculature: a basis for ischemic retinopathy. Cardiovasc. Res. 47, 489, 2000.<br />

119. Multicenter trial of cryotherapy for retinopathy of prematurity: natural history<br />

ROP: ocular outcome at 5(1/2) years in premature infants with birth weights less<br />

than 1251 g. Arch. Ophthalmol. 120, 595, 2002.<br />

120. Ashton, N., Ward, B., and Serpell, G.. Effect of oxygen on developing retinal<br />

vessels with particular reference to the problem of retrolental fibroplasia. Br. J.<br />

Ophthalmol. 38, 397, 1954.<br />

121. Stone, J. et al. Development of retinal vasculature is mediated by hypoxia-induced<br />

vascular endothelial growth factor (VEGF) expression by neuroglia. J. Neurosci.<br />

15, 4738, 1995.<br />

122. Carmeliet, P. Mechanisms of angiogenesis and arteriogenesis. Nat. Med. 6, 389,<br />

2000.


Oxidative Stress and Antioxidants in the Perinatal Period 91<br />

123. Helmlinger, G. et al. Formation of endothelial cell networks. Nature 405, 139,<br />

2000.<br />

124. Smith, L.E. Pathogenesis of retinopathy of prematurity. Semin. Neonatol. 8, 469,<br />

2003.<br />

125. Cunningham, S. et al. Transcutaneous oxygen levels in retinopathy of prematurity.<br />

Lancet 346, 1464, 1995.<br />

126. Saito, Y. et al. The progression of retinopathy of prematurity and fluctuation in<br />

blood gas tension. Graefes Arch. Clin. Exp. Ophthalmol. 231, 151, 1993.<br />

127. Supplemental therapeutic oxygen for prethreshold retinopathy of prematurity<br />

(STOP-ROP): a randomized, controlled trial. I. Primary outcomes. Pediatrics 105,<br />

295, 2000.<br />

128. Raju, T.N. et al. Vitamin E prophylaxis to reduce retinopathy of prematurity: a<br />

reappraisal of published trials. J. Pediatr. 131, 844, 1997.<br />

129. Anderson, G.H. Human milk feeding. Pediatr. Clin. North Am. 32, 335, 1985.<br />

130. Goldman, A.S., Goldblum, R.M., and Ganson, L.A. Anti-inflammatory systems<br />

in human milk. Adv. Exp. Med. Biol. 262, 69, 1990.<br />

131. Buescher, E.S. and McIlheran, S.M. Antioxidant properties of human colostrum.<br />

Pediatr. Res. 24, 14, 1988.<br />

132. L'Abbe, M.R. and Friel, J.K. Superoxide dismutase and glutathione peroxidase<br />

content of human milk from mothers of premature and full-term infants during<br />

the first 3 months of lactation. J. Pediatr. Gastroenterol. Nutr. 31, 270, 2000.<br />

133. Hamosh, M. Bioactive factors in human milk. Pediatr. Clin. North Am. 48, 69,<br />

2001.<br />

134. Friel, J.K. et al. Milk from mothers of both premature and full-term infants<br />

provides better antioxidant protection than does infant formula. Pediatr. Res. 51,<br />

612, 2002.<br />

135. Shoji, H. et al. Effects of human milk and spermine on hydrogen peroxide-induced<br />

oxidative damage in IEC-6 cells. J. Pediatr. Gastroenterol. Nutr. 41, 460, 2005.<br />

136. Zoeren-Grobben, D. et al. Postnatal changes in plasma chain-breaking antioxidants<br />

in healthy preterm infants fed formula and/or human milk. Am. J. Clin. Nutr. 60,<br />

900, 1994.<br />

137. Shoji, H. et al. Effect of human breast milk on urinary 8-hydroxy-2'-deoxyguanosine<br />

excretion in infants. Pediatr. Res. 53, 850, 2003.<br />

138. Shoji, H. et al. Suppressive effects of breast milk on oxidative DNA damage in<br />

very low birthweight infants. Arch. Dis. Child. Fetal Neonatal Ed. 89, F136, 2004.<br />

139. Lucas, A. and Cole, T.J. Breast milk and neonatal necrotising enterocolitis. Lancet<br />

336, 1519, 1990.<br />

140. Watkins, C.J., Leeder, S.R., and Corkhill, R.T. Relationship between breast and<br />

bottle feeding and respiratory illness in the first year of life. J. Epidemiol. Commun.<br />

Health 33, 180, 1979.<br />

141. Hylander, M.A. et al. Association of human milk feedings with a reduction in<br />

retinopathy of prematurity among very low birthweight infants. J. Perinatol. 21,<br />

356, 2001.<br />

142. Koletzko, B., Decsi, T., and Sawatzki, G.. Vitamin E status of low birthweight<br />

infants fed formula enriched with long-chain polyunsaturated fatty acids. Int. J.<br />

Vitam. Nutr. Res. 65, 101, 1995.<br />

143. Nair, V. et al. The chemistry of lipid peroxidation metabolites: crosslinking reactions<br />

of malondialdehyde. Lipids 21, 6, 1986.


92 Oxidative Stress and Inflammatory Mechanisms<br />

144. Decsi, T. and Koletzko, B. Growth, fatty acid composition of plasma lipid classes,<br />

and plasma retinol and alpha-tocopherol concentrations in full-term infants fed<br />

formula enriched with omega-6 and omega-3 long-chain polyunsaturated fatty<br />

acids. Acta Paediatr. 84, 725, 1995.<br />

145. Decsi, T., Burus, I., and Koletzko, B. Effects of dietary long-chain polyunsaturated<br />

fatty acids on plasma amino acids and indices of protein metabolism in infants:<br />

results from a randomized clinical trial. Ann. Nutr. Metab. 42, 195, 1998.<br />

146. Decsi, T., Burus, I., and Koletzko, B. Effects of dietary long-chain polyunsaturated<br />

fatty acids on plasma amino acids and indices of protein metabolism in infants:<br />

results from a randomized clinical trial. Ann. Nutr. Metab. 42, 195, 1998.


6<br />

CONTENTS<br />

Maternal Obesity,<br />

Glucose Intolerance,<br />

and Inflammation in<br />

Pregnancy<br />

Janet C. King<br />

Abstract ...............................................................................................................93<br />

Introduction .........................................................................................................94<br />

Obesity, Inflammation, and Insulin Resistance ..................................................94<br />

Pregnancy: An Inflammatory, Insulin-Resistant State........................................96<br />

Maternal Obesity and Inflammation...................................................................98<br />

Maternal Inflammation, Insulin Resistance, and Fetal Growth .......................100<br />

Interventions to Reduce Maternal Inflammation and Insulin Resistance ........101<br />

Conclusions .......................................................................................................103<br />

Acknowledgments .............................................................................................103<br />

References .........................................................................................................103<br />

ABSTRACT<br />

The prevalence of obesity among pregnant women is at an all-time high. Maternal<br />

obesity increases the risks of mortality and morbidity in the mother and baby.<br />

The risk of gestational diabetes, one of the most prevalent metabolic complications<br />

in pregnancy, is significantly greater among obese women. Emerging<br />

research suggests an association of obesity, inflammation, and insulin resistance<br />

in non-pregnant and pregnant individuals. Cytokines secreted by the placenta,<br />

such as tumor necrosis factor-α and leptin, may mediate that link in pregnancy.<br />

Studies of maternal body mass index (BMI), insulin resistance, and circulating<br />

levels of cytokines, i.e., tumor necrosis factor-α and C-reactive protein, show that<br />

maternal obesity is associated with increased levels of inflammatory markers and<br />

that elevated levels of these markers are linked to glucose intolerance in women.<br />

These metabolic adjustments may exacerbate fetal overgrowth and excessive<br />

93


94 Oxidative Stress and Inflammatory Mechanisms<br />

deposition of fat stores. Currently, obese pregnant women do not receive any<br />

guidance for reducing the risk of developing a pro-inflammatory, insulin-resistant<br />

state. Preliminary studies suggest that moderate physical activity and consuming<br />

a diet high in fiber and a higher proportion of polyunsaturated fatty acids may<br />

be beneficial.<br />

INTRODUCTION<br />

The prevalence of obesity among women of reproductive age has reached an alltime<br />

high. National survey data show that the number of women with body mass<br />

indexes (BMIs) greater than 30 averaged 31% in white women, 40% in Hispanic<br />

women, and 51% in black women in 1999 and 2000. 1 Obese women encounter<br />

more health problems during pregnancy. 1 Common disorders include pregnancyinduced<br />

hypertension, pre-eclampsia, large-for-gestational age (LGA) babies,<br />

need for cesarean or assisted deliveries, and post-delivery infections. The risk of<br />

gestational diabetes mellitus (GDM) is about three- to four-fold higher among<br />

obese compared to lean women. 2,3 The incidence of impaired glucose tolerance<br />

(IGT), i.e., one abnormal blood glucose value after an oral glucose load, is likely<br />

to be more frequent than GDM in obese women.<br />

The prevalence of GDM varies widely within and between populations.<br />

Although the U.S. national average is about 4%, 4 the prevalence is as high as<br />

16% in some high-risk populations. As with type 2 diabetes, the prevalence varies<br />

by race, ethnicity, and BMI. For example, in a study of 28,330 women from the<br />

Northern California Kaiser Permanente Medical Care Program, 7.4% of the Asian<br />

women, 5.6% of the Hispanic women, 4% of the African-American women, and<br />

3.9% of the white women developed GDM. 5 The rates among American Indian<br />

mothers were considerably higher: 15.1% among Zuni women and 10.4% among<br />

Navajo women. 4 Some of the differences due to race or ethnicity reflect the higher<br />

prevalence of obesity in certain populations. A study of 552 African-American<br />

women and 653 Latina women in Detroit showed that very obese Latina women<br />

(BMIs >35) demonstrated a 6.5-fold increased risk for developing GDM compared<br />

to normal weight Latina women; very obese African-American women had<br />

nearly a four-fold increased risk compared to normal weight women. 6 Although<br />

the risk of GDM increased with maternal body weights in both ethnic groups,<br />

the increase among Latina women was about 2.5 times greater than that in<br />

African-American women.<br />

OBESITY, INFLAMMATION, <strong>AND</strong> INSULIN RESISTANCE<br />

Obesity in non-pregnant adults is associated with subclinical inflammation and<br />

insulin resistance. 7 The inflammatory and insulin-resistant states arise from<br />

changes in cellular and molecular functions and metabolism when adipocytes<br />

become enlarged in obese individuals. Perlipin, a phosphoprotein on the surfaces<br />

of triglyceride droplets that acts as a gatekeeper preventing lipases from


Maternal Obesity, Glucose Intolerance, and Inflammation in Pregnancy 95<br />

Enlarged<br />

Adipocytes<br />

Macrophage<br />

Infiltration<br />

↑ TNF-α<br />

↑ IL-6<br />

↑ CRP<br />

INSULIN RESISTANCE<br />

↑ Lipolysis<br />

↑ Free Fatty Acids<br />

FIGURE 6.1 Relationship of obesity, inflammation, and insulin resistance. Enlarged adipocytes<br />

in the adipose tissues of obese individuals cause enhanced production of proinflammatory<br />

cytokines that increase rates of lipolysis, circulating levels of free fatty acids,<br />

and subsequent tissue insulin resistance.<br />

hydrolyzing triglycerides and releasing free fatty acids, declines when adipocytes<br />

become enlarged, leading to an increased release of free fatty acids. 7 In turn, free<br />

fatty acids promote insulin resistance by stimulating the phosphorylation of<br />

insulin-receptor substrate-1 (IRS-1) on its serine residues instead of the usual<br />

phosphorylation of tyrosine residues. 8 Once serine is phosphorylated, IRS-1<br />

becomes a poor substrate for the insulin receptor and insulin sensitivity declines.<br />

Cytokines, such as tumor necrosis factor-alpha (TNF-α) and interleukin-6 (IL-6)<br />

also stimulate IRS-1 phosphorylation on the serine residues.<br />

Obesity appears to predispose individuals to a pro-inflammatory state because<br />

enlarged fat stores promote the invasion of macrophages into adipose tissue to<br />

scavenge moribund adipocytes that tend to increase with obesity 7 (Figure 6.1).<br />

Thus, obese individuals often have increased circulating levels of cytokines such<br />

as TNF-α and IL-6. The cause of the increased number of moribund adipocytes<br />

in obese individuals is unknown, but one hypothesis is that clusters of adipocytes<br />

distant from capillary networks experience hypoxia before angiogenesis occurs. 9<br />

Cytokine secretion by macrophages also appears to be potently stimulated by<br />

acute-phase serum amyloid A; the decline in circulating levels of this protein<br />

after weight loss by obese individuals is associated with a reduction in circulating<br />

cytokine levels. 10<br />

The mild inflammatory state resulting from the increased secretion of cytokines<br />

by enlarged adipocytes contributes to the metabolic adjustments and insulin<br />

resistance associated with obesity. Some cytokines are thought to reduce adiponectin<br />

expression, 7 and serum adiponectin levels tend to fall with obesity. 11<br />

Adiponectin is a potent inhibitor of TNF-α-induced monocyte adhesion, which<br />

may explain in part the link between obesity and cardiovascular disease.<br />

TNF-α also increases adipocyte lipolysis, possibly by reducing perlipin 7 and<br />

thereby promoting the release of free fatty acids into circulation and reducing<br />

insulin sensitivity. IL-6 expression is also increased in the adipocytes of obese


96 Oxidative Stress and Inflammatory Mechanisms<br />

individuals with higher levels in visceral than in peripheral adipose tissue. 7 The<br />

elevated plasma IL-6 levels in obese individuals are also associated with an<br />

increase in lipolysis. Thus, both TNF-α and IL-6 enhance the breakdown of fats<br />

to free fatty acids in adipose tissue, increasing their delivery to liver and muscle<br />

and subsequently causing insulin resistance.<br />

Endogenous overproduction of cortisol or exogenous administration of cortisteroids<br />

generally causes weight gain with an increase in visceral fat stores<br />

compared with peripheral stores. However, obese people usually do not have<br />

higher levels of cortisol. 7 The activity of 11β-hydroxysteroid dehydrogenase type<br />

1, which converts inactive cortisol metabolites into active cortisol, is elevated in<br />

the adipose tissues of obese people. Higher rates of cortisol production in adipose<br />

tissue may contribute to hyperphagia, increased cytokine expression, gain in<br />

visceral adipose tissue, hyperlipidemia, and insulin resistance in obese individuals<br />

without any increases in circulating cortisol concentrations.<br />

Clearly, adipocytes are not merely storage reservoirs for fat, but they are also<br />

endocrine organs with multiple functions. Their metabolic function changes as<br />

they enlarge with increasing obesity. Enlarged adipocytes recruit macrophages,<br />

promote inflammation, and enhance the secretion of a variety of metabolites that<br />

predispose toward insulin resistance. An active area of research is identification<br />

of therapies that alter adipose tissue metabolic pathways leading to inflammation<br />

and insulin resistance.<br />

PREGNANCY: AN INFLAMMATORY, INSULIN-RESISTANT STATE<br />

Profound metabolic adjustments occur during pregnancy to assure an adequate<br />

nutrient supply is available to support fetal growth. Since glucose is the preferred<br />

fuel of the fetus, maternal metabolism is shifted toward a hyperglycemic state.<br />

This ensures facilitated glucose diffusion from maternal circulation across the<br />

placenta to the fetus. Maternal hyperglycemia is created by establishing an insulin-resistant<br />

state. Maternal insulin resistance increases throughout gestation,<br />

reaching a peak in late gestation when fetal fuel demands are the highest. 12 The<br />

rise in insulin resistance is ascribed to alterations in maternal cortisol levels and<br />

placental hormones (human placental lactogen, progesterone, and estrogen). 12<br />

However, the changes in insulin resistance have never been correlated with these<br />

hormonal changes in a prospective, longitudinal study. 13<br />

The recent evidence that adipokines such as TNF-α and leptin affect insulin<br />

sensitivity in non-pregnant individuals has led investigators to propose that similar<br />

mechanisms may occur in pregnancy. Pro-inflammatory cytokines may influence<br />

insulin metabolism in pregnant as well as non-pregnant women for several reasons.<br />

First, pregnant women with adequate food supplies gain fat during the first<br />

two trimesters. 1 The amounts gained vary from 0 to 10 kg and the average is<br />

about 3.5 kg. Expansion of the fat tissue and enlarged adipocytes may increase<br />

adipose tissue cytokine secretion and subsequent insulin resistance in pregnant<br />

women similar to that seen in non-pregnant individuals. Second, the placenta


Maternal Obesity, Glucose Intolerance, and Inflammation in Pregnancy 97<br />

expresses all known cytokines including TNF-α, IL-6, IL-10, leptin, resistin, and<br />

PAI-1. 14<br />

The physiological role of these placental cytokines is uncertain, but many of<br />

the same cytokines are produced by the placenta as are secreted by adipose tissue.<br />

Possibly, the high placental cytokine secretion assists in creating the maternal<br />

insulin-resistant state. Since maternal fat gain is very limited or non-existent in<br />

pregnant women with limited energy intakes, the placenta is likely to be more<br />

important for creating a maternal insulin-resistant state than is an increase in<br />

maternal adipose tissue.<br />

To test the hypothesis that placental cytokines play a role in modifying insulin<br />

sensitivity in pregnancy, Kirwan and co-workers measured circulating hormones,<br />

cytokines, and insulin resistance in 15 women (5 with GDM and 10 with normal<br />

glycemia) before pregnancy, at 12 to 14 weeks of gestation, and at 34 to 36 weeks<br />

of gestation. 13 Changes in insulin sensitivity were compared to placental hormones,<br />

cortisol, leptin, and TNF-α. TNF-α was more strongly associated with<br />

insulin sensitivity than any other hormone or cytokine measured; higher TNF-α<br />

levels were associated with lower insulin sensitivity (r = –0.69, p


98 Oxidative Stress and Inflammatory Mechanisms<br />

those observed in late pregnancy. 16 This confirmed that the levels of free fatty<br />

acids play an important role in regulating insulin sensitivity in pregnancy. Recent<br />

research suggests that the circulating levels of cytokines secreted by the placenta<br />

may mediate shifts in free fatty acid concentrations and subsequent insulin resistance.<br />

MATERNAL OBESITY <strong>AND</strong> INFLAMMATION<br />

Studies of inflammation and insulin resistance in pregnancy were performed in<br />

non-obese women. 13,17 Since obesity precipitates inflammatory responses, excessive<br />

free fatty acid release, and subsequent insulin-resistant states in non-pregnant<br />

individuals, it is reasonable to assume that inflammatory responses and insulin<br />

resistance would be enhanced in obese compared to lean pregnant women. Comparisons<br />

of metabolic adjustments in lean and obese pregnant women are limited,<br />

but the few studies done show that obese women rely more on fat oxidation as<br />

a source of energy in late pregnancy than do lean women. 18,19<br />

The increase in fat oxidation among the obese women was significantly<br />

correlated with serum leptin concentrations (r = 0.76, p


Maternal Obesity, Glucose Intolerance, and Inflammation in Pregnancy 99<br />

concentrations following oral glucose tolerance tests. 23 Insulin sensitivity declined<br />

by 34% in the non-obese group from 12 to 34 weeks and by 60% in the obese<br />

group; the change in the two groups did not achieve statistical significance (p =<br />

0.09). Insulin sensitivity did not differ between the two groups at 12 weeks of<br />

gestation, but it was significantly lower in the obese than in the non-obese at 34<br />

weeks (p


100 Oxidative Stress and Inflammatory Mechanisms<br />

MATERNAL INFLAMMATION, INSULIN RESISTANCE, <strong>AND</strong><br />

FETAL GROWTH<br />

Obese women tend to have big babies irrespective of the amount of weight they<br />

gain. 26 Also, maternal glucose metabolism and adiposity are highly correlated<br />

with fetal growth and body composition. 21,27 Although gestational age at birth is<br />

the strongest predictor of both birth weight and infant fat-free mass, maternal<br />

pregravid BMI is the strongest predictor of infant fat mass (r 2 = 0.066), explaining<br />

about 7% of the variance in body fat of newborns.<br />

Maternal diabetes is a strong predictor of fetal overgrowth, with diabetic<br />

women facing a three-fold higher risk of having an LGA baby than obese women.<br />

Four times as many LGA infants are born to obese women than to diabetic women<br />

because there are many more obese women than diabetic women. Also, diabetic<br />

women usually are carefully monitored by their health care providers during<br />

pregnancy whereas no special care is provided to obese women. The additional<br />

care given to diabetic women undoubtedly reduces the number of LGA babies<br />

born to them. Recent studies from both North America and Europe report<br />

increases in mean birth weight over the past 30 years. 21 This rise in birth weight<br />

is likely related to the increasing incidence of maternal obesity. In fact, Catalano<br />

and his co-workers in Cleveland, Ohio found that a mean increase in birth weight<br />

of 116 g over the past 30 years was more strongly correlated with maternal weight<br />

at delivery than any other maternal characteristic.<br />

Is this rise in birth weight linked to an increased maternal subclinical inflammation<br />

and insulin resistance associated with obesity? Preliminary data suggest<br />

that it may be. 28 For example, Radaelli and co-workers 29 measured circulating<br />

maternal and fetal cytokines and growth factors in three mother–infant cohorts<br />

divided into tertiles according to neonatal body fat. The only fetal factor associated<br />

with neonatal body fat was leptin (p


Maternal Obesity, Glucose Intolerance, and Inflammation in Pregnancy 101<br />

INTERVENTIONS TO REDUCE MATERNAL INFLAMMATION <strong>AND</strong><br />

INSULIN RESISTANCE<br />

Studies in non-pregnant obese individuals provided evidence that changes in<br />

lifestyle (weight reduction, changes in dietary fat and fiber, and increased physical<br />

activity) reduced the risks of type 2 diabetes mellitus in obese individuals. 31<br />

Weight loss is a very effective way to reduce circulating cytokine levels and<br />

fasting insulin concentrations in obese women 32 but weight loss is not indicated<br />

for pregnant women. However, modest reductions in the total amounts gained by<br />

obese women are appropriate. The Institute of Medicine recommends that obese<br />

women gain at least 15 lb during gestation; normal weight women are advised<br />

to gain 25 to 35 lb. 33<br />

Physical activity is an effective intervention for reducing the risk of type 2<br />

diabetes and associated metabolic anomalies such as insulin resistance, oxidative<br />

stress, and dyslipidemia. 34 Physical activity activates the AMP-activated protein<br />

kinase (AMPK) enzyme, which increases glucose transport into the muscle,<br />

enhances fat oxidation, and reduces insulin resistance. 7 . Exercise, even intermittently,<br />

reduces the risk of GDM among obese women with BMIs >33 by nearly<br />

two-fold. 35 Women who exercise throughout pregnancy (i.e., perform endurance<br />

exercises ≥4 times/week) gain significantly less fat and had significantly lower<br />

increases in TNF-α and leptin during gestation. 36 The changes in leptin, but not<br />

TNF-α, were correlated with reduced fat mass in physically active women.<br />

Possibly, the differences in TNF-α levels reflect the exercise-induced reductions<br />

in insulin resistance whereas the leptin changes are more closely linked to fat<br />

accretion. Nevertheless, moderate physical activity during pregnancy may be an<br />

effective way to reduce subclinical inflammation and insulin resistance during<br />

pregnancy.<br />

Changes in the types of dietary fat and carbohydrates consumed may be other<br />

ways to reduce maternal inflammation and insulin resistance. Decreases in total<br />

fat and saturated fat intakes and increases in dietary fiber are recommended for<br />

reducing the risk of type 2 diabetes mellitus in non-pregnant adults. 37 Results<br />

from the limited number of studies of pregnant women suggest that similar<br />

changes in dietary fats and carbohydrates also effectively reduced the risks of<br />

glucose intolerance. Clapp randomized 12 pregnant women to a low or high<br />

glycemic index diet prior to conception. 38 The amounts of carbohydrate consumed<br />

were similar in the two groups: 56% of the total energy.<br />

During pregnancy, the women on the low glycemic diet showed no significant<br />

changes in their glycemic responses to mixed meals whereas the women on the<br />

high glycemic diet experienced 190% increases in their responses compared to<br />

pre-pregnancy values. These findings are similar to those reported by Fraser et<br />

al. 39 who showed that a high fiber diet reduced the post-prandial response to a<br />

meal in comparison to low fiber intake. Bronstein and co-workers have also shown<br />

that reducing the glycemic index of a test meal lowered post-prandial glucose<br />

and insulin responses in both lean and obese women studied in the third trimester 40<br />

(Figure 6.3). These findings suggest that the maternal insulin-resistant state and


102 Oxidative Stress and Inflammatory Mechanisms<br />

mg/mLxhrs<br />

uU/mLxhrs<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

900<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

Insulin Area Under Curve<br />

High GL Lower GL<br />

Glucose Area Under Curve<br />

High GL Lower GL<br />

Obese<br />

Non-obese<br />

FIGURE 6.3 Serum insulin and glucose responses to lower glycemic test meals in lean<br />

and obese pregnant women. The areas under the curve for serum insulin and glucose were<br />

measured over a 2-hour period in 6 lean and 8 obese pregnant women studied at 32 to 36<br />

weeks of gestation. 40<br />

hyperglycemic responses to meals reflect the intake of a Westernized, low fiber,<br />

high glycemic diet rather than a typical metabolic response to pregnancy.<br />

Studies of non-pregnant, obese individuals also suggest that increasing the<br />

ratio of polyunsaturated to saturated fats in the diet may reduce the risk of<br />

developing metabolic syndrome and its complications as early as adolescence. 41<br />

Similar findings have been reported for pregnant women. In a study of 171<br />

pregnant Chinese women with or without impaired glucose tolerance, the type<br />

of dietary fat predicted impaired glucose tolerance and GDM. 42 In a logistic<br />

regression analysis, increased body weight, decreased polyunsaturated fat intake,<br />

and a low dietary polyunsaturated-to-saturated fat ratio independently predicted<br />

glucose intolerance. Bo and co-workers also found that glucose intolerance in<br />

pregnant women without conventional risk factors (i.e., family history, age, and


Maternal Obesity, Glucose Intolerance, and Inflammation in Pregnancy 103<br />

BMI) was related to the percent of saturated and polyunsaturated fats in the diet<br />

with high intakes of saturated fat increasing the risk and high intakes of polyunsaturated<br />

fat decreasing the risk. 43<br />

The conventional dietary treatment of women with GDM is to reduce the<br />

amount of dietary carbohydrate and increase slightly the amount of fat. The<br />

preliminary data reviewed here suggest that it is more important to consider the<br />

types of carbohydrates and fats rather than the amounts. Increasing dietary fiber<br />

(or lowering the glycemic index) and the proportion of polyunsaturated fatty acids<br />

may be effective interventions for reducing inflammation and insulin resistance<br />

in pregnancy.<br />

CONCLUSIONS<br />

Placental cytokine secretion induces a pro-inflammatory state during pregnancy.<br />

This metabolic state is exacerbated by additional cytokines secreted by enlarged<br />

adipocytes in obese pregnant women. The inflammatory state leads to an increase<br />

in circulating free fatty acids and an insulin-resistant state. This pro-inflammatory,<br />

insulin-resistant state in a mother may affect placental function and fetal growth<br />

and development. Accumulating evidence suggests that the fetal fuel supply is<br />

enhanced, leading to accelerated growth and excessive fetal fat stores. Currently,<br />

obese pregnant women are not given any specific advice for reducing the proinflammatory<br />

response and insulin resistance. However, studies in both nonpregnant<br />

and pregnant adults suggest that moderate physical activity and<br />

increased intakes of dietary fiber and the proportion of polyunsaturated fatty acids<br />

attenuate these metabolic abnormalities and thereby potentially improve pregnancy<br />

outcomes.<br />

ACKNOWLEDGMENTS<br />

The author acknowledges the contributions of her research collaborators to the<br />

original, unpublished research findings included in this paper: Jessica DeHaene,<br />

Dina El Kady, James L. Graham, Peter J. Havel, Liza Kunz, Meredith Milet,<br />

Ratna Mukherjea, Kimber L. Stanhope, and Leslie R. Woodhouse.<br />

REFERENCES<br />

1. J.C. King. Maternal obesity, metabolism, and pregnancy outcomes. Annu Rev Nutr<br />

2006.<br />

2. L.C. Castro and Avina R.L. Maternal obesity and pregnancy outcomes. Curr Opin<br />

Obstet Gynecol 2002; 14: 601.<br />

3. K.R. Andreasen, Andersen M.L., and Schantz A.L. Obesity and pregnancy. Acta<br />

Obstet Gynecol Scand 2004; 83: 1022.


104 Oxidative Stress and Inflammatory Mechanisms<br />

4. G.L.A. Beckles and Thompson-Reid P.E. Diabetes and women's health across life<br />

states: a public health perspective (translation). Centers for Disease Control and<br />

Prevention, National Center for Chronic Disease Prevention and Health Promotion,<br />

Atlanta, 2001.<br />

5. A. Ferrara et al. Prevalence of gestational diabetes mellitus detected by the<br />

National Diabetes Data Group or the Carpenter and Coustan plasma glucose<br />

thresholds. Diabetes Care 2002; 25: 1625.<br />

6. E.C. Kieffer et al. Obesity and gestational diabetes among African-American<br />

women and Latinas in Detroit: implications for disparities in women's health. J<br />

Am Med Womens Assn 2001; 56: 181.<br />

7. A.S. Greenberg and Obin M.S. Obesity and the role of adipose tissue in inflammation<br />

and metabolism. Am J Clin Nutr 2006; 83: 461S.<br />

8. G.S. Hotamisligil. Inflammatory pathways and insulin action. Int J Obes 2003;<br />

27: 553.<br />

9. P. Trayhurn. Endocrine and signalling role of adipose tissue: new perspectives on<br />

fat. Acta Physiol Scand 2005; 184: 285.<br />

10. R.Z. Yang et al. Acute-phase serum amyloid A: an inflammatory adipokine and<br />

potential link between obesity and its metabolic complications. Plos Med 2006;<br />

3: 884.<br />

11. P.J. Havel. Control of energy homeostasis and insulin action by adipocyte hormones:<br />

leptin, acylation stimulating protein, and adiponectin production. Curr<br />

Opin Lipidol 2002; 13: 51.<br />

12. J.L. Kitzmiller. The endocrine pancreas and maternal metabolism, in D. Tulchinsky<br />

and Ryan K.J., Eds, Maternal–Fetal Endocrinology, W.B. Saunders, Philadelphia,<br />

1980, p. 58.<br />

13. J.P. Kirwan et al. TNF-α is a predictor of insulin resistance in human pregnancy.<br />

Diabetes 2002; 51: 2207.<br />

14. S. Hauguel-de Mouzon and Guerre-Millo M. The placenta cytokine network and<br />

inflammatory signals. Placenta 2005.<br />

15. R.L. Phelps, Metzger B.E., and Freinkel N. Carbohydrate metabolism in pregnancy.<br />

XVII. Diurnal profiles of plasma glucose, insulin, free fatty acids, triglycerides,<br />

cholesterol, and individual amino acids in late normal pregnancy. Am J Obstet<br />

Gynecol 1981; 140: 730.<br />

16. E. Sivan et al. Free fatty acids and insulin resistance during pregnancy. J Clin<br />

Endocrinol Metab 1998; 83: 2338.<br />

17. T. Radaelli et al. Gestational diabetes induces placental genes for chronic stress<br />

and inflammatory pathways. Diabetes 2003; 52: 2951.<br />

18. P.M. Catalano et al. Carbohydrate metabolism during pregnancy in control subjects<br />

and women with gestational diabetes. Am J Physiol 1993; 264: E60.<br />

19. N.C. Okereke et al. Longitudinal changes in energy expenditure and body composition<br />

in obese women with normal and impaired glucose tolerance. Am J<br />

Physiol Endocrinol Metab 2004; 287: E472.<br />

20. N.C. Okereke et al. Longitudinal changes in energy expenditure and body composition<br />

in obese women with normal and impaired glucose tolerance. Am J<br />

Physiol 2004, publication.<br />

21. P. Catalano and Ehrenberg H. The short- and long-term implications of maternal<br />

obesity on the mother and her offspring. Bjog 2006.<br />

22. K.E. Lang. Maternal metabolic and pregnancy outcomes in obese and non-obese<br />

women. Unpublished data, University of California, Davis, 2006.


Maternal Obesity, Glucose Intolerance, and Inflammation in Pregnancy 105<br />

23. M. Matsuda and DeFronzo R.A. Insulin sensitivity indices obtained from oral<br />

glucose tolerance testing: comparison with the euglycemic insulin clamp. Diabetes<br />

Care 1999; 22: 1462.<br />

24. L. Kunz. Circulating levels of cytokines during the third trimester of pregnancy.<br />

Unpublished data, University of California, Davis, 2006.<br />

25. J.E. Ramsay et al. Maternal obesity is associated with dysregulation of metabolic,<br />

vascular, and inflammatory pathways. J Clin Endocrinol Metab 2002; 87: 4231.<br />

26. B.F. Abrams and Laros R.K., Jr. Pre-pregnancy weight, weight gain, and birth<br />

weight. Am J Obstet Gynecol 1986; 154: 503.<br />

27. P.M. Catalano et al. Increased fetal adiposity: a very sensitivie marker of abnormal<br />

in utero development. Am J Obstet Gynecol 2003; 189: 1698.<br />

28. P.M. Catalano. Obesity and pregnancy: the propagation of a viscous cycle?<br />

(editorial). J Clin Endocrinol Metab 2003; 88: 3505.<br />

29. T. Radaelli et al. Maternal interleukin-6: marker of fetal growth and adiposity. J<br />

Soc Gynecol Investig 2006; 13: 53.<br />

30. K.W. Huggins, Boileau A.C., and Hui D.Y. Portection against diet-induced obesity<br />

and obesity-related insulin resistance in group 1B PLA-2-deficient mice. Am J<br />

Physiol Endocrinol Metab 2002; 283: E994.<br />

31. J. Tuomilehto et al. Prevention of type 2 diabetes mellitus by changes in lifestyle<br />

among subjects with impaired glucose tolerance. New Engl J Med 2001; 344:<br />

1343.<br />

32. P Ziccardi et al. Reduction of inflammatory cytokine concentrations and improvement<br />

of endothelial functions in obese women after weight loss over one year.<br />

Circulation 2002; 105: 804.<br />

33. Food and Nutrition Board Institute of Medicine. Nutrition during Pregnancy. Part<br />

I. Weight Gain. Part II. Nutrient Supplements. National Academy Press, Washington,<br />

1990.<br />

34. J.C. Dempsey et al. Prospective study of gestational diabetes mellitus risk in<br />

relation to maternal recreational physical activity before and during pregnancy.<br />

Am J Epidemiol 2004; 159: 663.<br />

35. T.D. Dye et al. Physical activity, obesity, and diabetes in pregnancy. Am J Epidemiol<br />

1997; 146: 961.<br />

36. J.F. Clapp, 3rd and Kiess W. Effects of pregnancy and exercise on concentrations<br />

of the metabolic markers tumor necrosis factor alpha and leptin. Am J Obstet<br />

Gynecol 2000; 182: 300.<br />

37. U.S. Department of Health and Human Services and U.S. Department of Agriculture,<br />

Dietary Guidelines for Americans, U.S. Government Printing Office,<br />

Washington, 2005.<br />

38. J.F. Clapp, 3rd. Effect of dietary carbohydrate on the glucose and insulin response<br />

to mixed caloric intake and exercise in both nonpregnant and pregnant women.<br />

Diab Care 1998; 21: B107.<br />

39. R.B. Fraser, Ford F.A., and Lawrence G.F. Insulin sensitivity in third trimester<br />

pregnancy: a randomized study of dietary effects. Br J Obstet Gynaecol 1988; 95:<br />

223.<br />

40. M.N. Bronstein, Mak R.P., and King J.C. The thermic effect of food in normal<br />

weight and overweight pregnant women. Br J Nutr 1995; 75: 261.<br />

41. C. Klein-Platat et al. Plasma fatty acid composition is associated with the metabolic<br />

syndrome and low-grade inflammation in overweight adolescents. Am J Clin<br />

Nutr 2005; 82: 1178.


106 Oxidative Stress and Inflammatory Mechanisms<br />

42. Y. Wang et al. Dietary variables and glucose tolerance in pregnancy. Diabetes<br />

Care 2000; 23: 460.<br />

43. S. Bo et al. Dietary fat and gestational hyperglycaemia. Diabetologia 2001; 44:<br />

972.


7<br />

CONTENTS<br />

Obesity, Nutrigenomics,<br />

Metabolic Syndrome,<br />

and Type 2 Diabetes<br />

David Heber<br />

Introduction .......................................................................................................107<br />

Obesity Epidemic and Its Solutions .................................................................108<br />

Type 2 Diabetes Mellitus and Obesity: “Diabesity”........................................111<br />

Metabolic Syndrome: Pre-Diabetes or Cardiometabolic Risk? .......................111<br />

Nutrigenetics of Type 2 Diabetes Mellitus.......................................................113<br />

Beta Cell Failure and Type 2 Diabetes Mellitus..............................................115<br />

Adipokines ........................................................................................................116<br />

Conclusion.........................................................................................................118<br />

References .........................................................................................................118<br />

INTRODUCTION<br />

The traditional medical paradigm for understanding diabetes mellitus has undergone<br />

a major shift in the past two decades with the discovery of the many<br />

immunological functions of fat cells, especially those located in the mesenteric<br />

or visceral fat of the abdomen. This shift in thinking occurred in the molecular–genetic<br />

era of the last decade, beginning with the discovery of leptin.<br />

Once it was realized that leptin is also a cytokine that stimulates angiogenesis,<br />

the many connections between obesity, inflammation, and diabetes began to<br />

emerge. A large number of cytokines and chemokines originating in fat cells were<br />

subsequently discovered. Recently, our group has demonstrated that a small but<br />

significant 5% weight loss could lead to a much larger decrease in the levels of<br />

circulating C-reactive protein, a biomarker of inflammation. This connection of<br />

inflammation and obesity, leading in genetically susceptible individuals to diabetes<br />

mellitus type 2, provides a unifying pathophysiological mechanism for<br />

many of the co-morbid diseases associated with diabetes and the metabolic<br />

syndrome including cardiovascular disease, renal disease, liver disease, and<br />

hypertension.<br />

107


108 Oxidative Stress and Inflammatory Mechanisms<br />

OBESITY EPIDEMIC <strong>AND</strong> ITS SOLUTIONS<br />

Obesity is defined as excess body fat, but the difficulties inherent in measuring<br />

body fat in tens of thousands of individuals have resulted in the substitution of<br />

a surrogate measure called body mass index (BMI), defined as weight in kilograms<br />

over height in meters squared. This is a regression equation that approximates<br />

excess body fat in large populations, but can be significantly erroneous in<br />

individuals such as athletes, whose excess weight is due to muscle, or in young<br />

sedentary women consuming inadequate dietary protein who may have low body<br />

weights with abnormally high percentages of body fat.<br />

Practical methods such as bioelectrical impedance can be used to assess body<br />

fat in individuals in clinical settings, but the BMI has proven invaluable in<br />

population studies. Overweight, defined as a BMI of at least 25 but less than 30<br />

kg/m 2 , and obesity (used here to denote greater overweight), defined as a BMI<br />

of 30 kg/m 2 or greater, are major contributors to morbidity and mortality in the<br />

United States today. The risk for some of our most devastating diseases, especially<br />

cardiovascular disease and diabetes mellitus, is significantly correlated with an<br />

individual’s BMI 1 and a huge portion of the country's healthcare resources are<br />

used to treat the consequences of overweight and obesity. The United States is<br />

in the midst of an explosive epidemic of obesity. Although the early maps from<br />

1985 to 1990 present all states in shades of light blue, indicating obesity prevalences<br />

of less than 15%, the latest maps are almost entirely orange (20 to 24%<br />

prevalence), with an ominous streak of red (25% or more) arcing from Texas to<br />

Michigan through Appalachia. 2 (Maps are available at http://www.cdc.gov/nccdphp/dnpa/obesity/trend/maps/.)<br />

Weight gain is primarily a matter of energy imbalance — energy intake that<br />

is greater than energy output. This imbalance is influenced by genetic, metabolic,<br />

behavioral, and environmental factors. Only the latter two factors could have<br />

changed enough in the last 20 years to cause the obesity epidemic, resulting in<br />

both an increase in average energy intake and a decrease in average energy<br />

expenditure. Energy intake has increased in the United States due to a greater<br />

availability of highly palatable and energy-dense foods, larger portions becoming<br />

the norm, more food consumed outside the home (where less heed is paid to<br />

nutrition), and greater consumption of high-sugar beverages and high-fat processed<br />

and fast foods.<br />

Over the past 20 years, nutrition researchers and food scientists over-emphasized<br />

reducing fat consumption with less regard for total calories, resulting in an<br />

ineffective low-fat food campaign while the incidence of obesity continued to<br />

increase. A significant decrease in energy output has been associated with<br />

increased hours of television viewing, increased numbers of individuals in sedentary<br />

jobs, and the increased prevalence of labor-saving devices such as elevators,<br />

automobiles, and remote controls. Some experts hold that the current obesity<br />

epidemic is caused by relatively small but chronic energy imbalances. Because<br />

a pound of body fat represents 3500 stored calories, a chronic energy imbalance<br />

— intake over output — of only 100 calories per day will cause a weight gain


Obesity, Nutrigenomics, Metabolic Syndrome, and Type 2 Diabetes 109<br />

TABLE 7.1<br />

Patient Assessment: Factors to Consider in Developing Weight Loss<br />

Strategies<br />

Factor Guidelines<br />

Body mass BMI 25.0 to 29.9 kg/m<br />

index (BMI)<br />

2 (overweight)<br />

BMI ≥30 kg/m2 (obese)<br />

Waist<br />

Men: >102 cm (40 inches)<br />

circumference Women: >88 cm (35 inches)<br />

Risk status High:<br />

Presence of coronary heart disease, atherosclerosis, type 2 diabetes, sleep<br />

apnea, OR any three of the following:<br />

Cigarette smoking<br />

Hypertension (systolic blood pressure ≥140 mm Hg or diastolic blood pressure<br />

≥90 mm Hg)<br />

High LDL cholesterol (≥160 mg/dL)<br />

Low HDL cholesterol (200 mg/dL)<br />

Patient<br />

Reasons and motivation for weight reduction<br />

motivation History of successful and unsuccessful weight loss attempts<br />

Support (family, friends, co-workers)<br />

Patient's understanding of causes and health risks of obesity<br />

Attitude toward and capacity for physical activity<br />

Time available for weight loss intervention<br />

Financial considerations<br />

Adapted from NIH/NHLBI clinical guidelines on overweight and obesity. Body mass index (kg/m 2 )<br />

can be calculated from pounds and inches using the formula: BMI (kg/m 2 ) = [weight<br />

(pounds)/height (inches) 2 ] × 704.5. An interactive BMI calculator and BMI chart are included in<br />

the NIH/NHLBI report.<br />

of 10 pounds a year. The only way to restore energy balance is to reduce energy<br />

intake and/or increase energy output.<br />

The assessment process, as described in the National Institutes of Health/<br />

National Heart, Lung, and Blood Institute report titled Clinical Guidelines on the<br />

Identification, Evaluation, and Treatment of Overweight and Obesity in Adults,<br />

involves considering a patient's BMI, waist circumference, motivation to lose<br />

weight, and overall risk status (Table 7.1).<br />

Weight reduction has well documented health benefits and can usually be<br />

achieved with a loss of 1 to 2 pounds per week (0.5 to 1.0 pounds for lower


110 Oxidative Stress and Inflammatory Mechanisms<br />

starting weights), requiring an energy deficit of 500 to 1000 calories per day for<br />

the obese or 300 to 500 calories per day for the overweight. Achieving this energy<br />

deficit requires a three-part strategy that includes dietary therapy, physical activity,<br />

and behavior therapy.<br />

Dietary therapy may be the most controversial of these elements. Patients<br />

tend to diet to achieve weight loss and then stop when their goals are reached.<br />

However, obesity is a chronic disease that cannot be cured and must be controlled<br />

through permanent lifestyle changes. Indications of what comprises successful<br />

diet strategies have been gleaned from the National Weight Control Registry<br />

(NWCR), which is following more than 4500 adults who were able to maintain<br />

weight losses of at least 30 pounds for at least a year. The typical participant<br />

(with an average weight loss of 60 pounds maintained for 5 years) made a<br />

permanent change to a low-calorie diet, what is called “chronic restrained eating.” 4<br />

There are many possible means to attaining a state of chronic restrained<br />

eating. Thus, it makes little difference which form of diet one chooses to follow,<br />

whether it involves restricted carbohydrates (e.g., Atkins), macronutrient balance<br />

(e.g., Zone), restricted fat (e.g., Ornish), or simply restricted portion sizes and<br />

calories (e.g., Weight Watchers); what matters primarily is whether one is able<br />

to adhere to the diet. 5 Once a weight goal is achieved, there is general agreement<br />

that an optimally healthy diet consisting of seven servings per day of colorful<br />

fruits and vegetables, whole grains, low fat protein sources, and limited amounts<br />

of refined carbohydrates should be consumed. Significant amounts of research<br />

also suggest that structured diet plans using meal replacements such as high<br />

protein shakes or frozen meals are helpful adjuncts to weight loss. The most<br />

important requirements for an effective diet are that it (1) establishes a calorie<br />

deficit, (2) is healthy, and (3) fits one’s lifestyle.<br />

The next component of a successful weight loss strategy is physical activity.<br />

In fact, a daily expenditure of 300 or more calories through physical activity may<br />

be the most important factor for maintaining weight loss according to the NWCR.<br />

To accomplish this, a formerly obese person would need to perform 60 to 90<br />

minutes per day of moderate-intensity physical activity, equivalent to walking at<br />

3.5 to 4.0 miles per hour. Lesser amounts are recommended to prevent excess<br />

weight gain or reduce the risk of chronic disease.<br />

A popular and simple program that started in Japan is the 10,000 steps per<br />

day idea, which provides a concrete goal. Since the goal can be worked on<br />

throughout the day, this strategy encourages cumulative episodes of physical<br />

activity. A goal counted in daily steps has been shown to induce more total<br />

walking than a goal measured in, say, hours of brisk walking. 6<br />

Abdominal obesity due to excess visceral fat is associated with an increased<br />

risk of developing cardiovascular disease. 1,7 Moreover, excess visceral fat is<br />

linked to an increased risk of metabolic syndrome, which includes a greater risk<br />

of developing type 2 diabetes mellitus 3 with its associated cardiometabolic<br />

disorders. 4


Obesity, Nutrigenomics, Metabolic Syndrome, and Type 2 Diabetes 111<br />

TYPE 2 DIABETES MELLITUS <strong>AND</strong> OBESITY: “DIABESITY”<br />

The focus of physicians treating diabetes has been strictly on glucose control for<br />

much of the last century. In fact, the term diabetes mellitus derives from the Latin<br />

meaning “sweet urine.” In Ayurvedic medicine, the term for diabetes<br />

(madhumeda) translates as “one whose urine attracts ants.” This glucocentric<br />

model of diabetes evolved from observations in the 1920s that surgical removal<br />

of the pancreas in a dog led to diabetes mellitus. Insulin injections could then<br />

restore glucose metabolism.<br />

In juvenile or type 1 diabetes mellitus patients with autoimmune destruction<br />

of the pancreas early in life, insulin treatment could delay or prevent many of<br />

the complications of this terrible disease such as blindness, renal failure, and<br />

need for limb amputations. However, this form of diabetes is much less common<br />

today, accounting for less than 5% of all diabetes cases. Today, over 95% of all<br />

diabetes is type 2 and it occurs in children and adolescents as well as adults. In<br />

an individual with a BMI of 30, the risk of diabetes type 2 is increased 60- to<br />

80-fold in comparison to lean individuals. In contrast, the risk for heart disease<br />

is only 4- to 6-fold increased at a BMI of 30 (see Figure 7.1). The association<br />

of diabetes type 2 with obesity goes beyond typical risk factors and justifies<br />

naming the disease diabesity. While 10% of type 2 diabetes patients are said to<br />

be lean, this assessment is based on weight and not body composition; many of<br />

these individuals may have excess abdominal fat.<br />

As type 2 diabetes mellitus gained recognition in the 1970s, its etiology<br />

remained poorly understood. The idea that insulin, by controlling complications,<br />

was central to therapy was simply applied to type 2 diabetes mellitus as it had<br />

been to type 1, with the assumption that the results would be the same. However,<br />

insulin treatment failed to reduce cardiovascular mortality.<br />

The late onset of diabetes mellitus type 2, usually in the fourth to sixth decade<br />

of life, was attributed to aging of the beta cell within the pancreas which secretes<br />

insulin. The pathophysiologic progression of type 2 diabetes mellitus was first<br />

described by Drenick and Johnson. Over a 2- to 10-year period, their model<br />

predicted evolution from hyperinsulinemia with euglycemia to a condition of<br />

insulin deficiency and hyperglycemia as the beta cell was exhausted. Dr. Peter<br />

Butler recently described a likely cause of beta cell exhaustion by uncovering<br />

the role of an amyloid protein called insulin-associated polypeptide or IAPP.<br />

METABOLIC SYNDROME: PRE-DIABETES OR<br />

CARDIOMETABOLIC RISK?<br />

A particular cluster of risk factors that seems especially coherent and predictive<br />

of atherosclerotic cardiovascular disease (ASCVD) has been called the metabolic<br />

syndrome. It comprises abdominal obesity, atherogenic dyslipidemia, hypertension,<br />

and elevated plasma glucose, along with the pro-thrombotic and pro-inflammatory<br />

states. Identification of this syndrome has proven useful in unifying


112 Oxidative Stress and Inflammatory Mechanisms<br />

Relative Risk<br />

(a)<br />

Relative Risk<br />

(b)<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Women<br />

Type 2 diabetes<br />

Cholelithiasis<br />

Hypertension<br />

Coronary heart disease<br />

21 22 23 24 25 26 27 28 29 30<br />

Body-Mass Index<br />

Men<br />

Type 2 diabetes<br />

Cholelithiasis<br />

Hypertension<br />

Coronary heart disease<br />

21 22 23 24 25 26 27 28 29 30<br />

Body-Mass Index<br />

FIGURE 7.1 Relation of body mass index (BMI) and relative risk of diabetes, cholelithiasis,<br />

hypertension, and coronary heart disease. (Source: From Willett, W.C. et al. New<br />

Engl J Med 341, 427, 1999.)<br />

medical approaches to the previously disparate realms of diabetes and cardiovascular<br />

disease. 8<br />

The National Heart, Lung, and Blood Institute in conjunction with the American<br />

Heart Association recently published updated guidelines for the diagnosis<br />

of metabolic syndrome 9 (Table 7.2). Note that elevated low-density lipoprotein<br />

cholesterol (LDL-C) is not listed as a component; this is because LDL-C levels<br />

are not generally divergent from normal levels in metabolic syndrome. However,<br />

LDL particles are seen to be smaller and denser in the syndrome, so that small,<br />

dense LDL (sdLDL) particle size is an experimental observation in metabolic<br />

syndrome but is not practical to assess clinically. The unifying framework for the<br />

relationships of obesity, metabolic syndrome, type 2 diabetes, and cardiovascular<br />

disease is hypothesized to be the release of free fatty acids (FFAs) preferentially


Obesity, Nutrigenomics, Metabolic Syndrome, and Type 2 Diabetes 113<br />

TABLE 7.2<br />

Criteria for Diagnosing Metabolic Syndrome<br />

Three or more of the following characteristics define metabolic syndrome:<br />

1. Abdominal obesity: waist circumference >102 cm in men and >88 cm in women<br />

2. Hypertriglyceridemia: ≥150 mg/dL (1.69 mmol/L)<br />

3. Low high-density lipoprotein (HDL) cholesterol:


114 Oxidative Stress and Inflammatory Mechanisms<br />

the past 100,000 years. It is likely that many forms of diabetes mellitus type 2<br />

are associated with numerous genetic polymorphisms concerning both inflammation<br />

and the adaptation to starvation.<br />

As of October 2005, 176 human obesity cases due to single-gene mutations<br />

in 11 different genes have been reported, 50 loci related to Mendelian syndromes<br />

relevant to human obesity have been mapped to a genomic region, and causal<br />

genes or strong candidates have been identified for most of these syndromes. 13<br />

There are 244 genes that, when mutated or expressed as transgenes in mice,<br />

result in phenotypes that affect body weight and adiposity. The number of studies<br />

reporting associations between DNA sequence variation in specific genes and<br />

obesity phenotypes has also increased considerably, with 426 findings of positive<br />

associations with 127 candidate genes. The nutrigenetics of diabetes mellitus type<br />

2 and obesity will most likely be closely related, if not identical, in many<br />

individuals with the difference being that environmental influences uncover<br />

underlying type 2 diabetes mellitus.<br />

The famous example of the Pima Indians in the Southwest United States<br />

points out the influence of environmental factors. Living only a few hundred<br />

miles apart, members of the same tribe were separated by the United States–Mexican<br />

border but were genetically similar. Members of the tribe living in the United<br />

States were forced onto a reservation in Arizona where they could not pursue<br />

their agricultural lifestyle. As a result, the age- and sex-adjusted prevalence of<br />

type 2 diabetes in the Mexican Pima Indians (6.9%) was less than one-fifth that<br />

in the U.S. Pima Indians (38%) and similar to that of non-Pima Mexicans (2.6%).<br />

The prevalence of obesity was similar in the Mexican Pima Indians (7% in men<br />

and 20% in women) and non-Pima Mexicans (9% in men and 27% in women)<br />

but was much lower than in the U.S. Pima Indians. 14 Other examples can be drawn<br />

both from immigrants to the U.S. or internationally in countries such as China,<br />

Japan, Korea, and Thailand where Western diets are rapidly replacing traditional<br />

Asian diets, leading to increases in obesity and diabetes.<br />

While the treatment strategies for this vast complex of medical problems<br />

traditionally includes treatment of the complications (hypertension, dyslipidemia,<br />

and type 2 diabetes), or management of the ensuing coronary heart disease risk,<br />

it is a challenge to incorporate prevention and reduction of the “hypertriglyceridemic<br />

waist” into standard medical practice.<br />

In patients who have already developed type 2 diabetes mellitus, a knowledge<br />

of proper blood glucose monitoring along with the ability to assess emotional<br />

status and garner adequate social support augment the efforts to increase physical<br />

activity and modify nutritional habits. Among the drugs used that can complement<br />

lifestyle change, metformin, which is off patent currently, is the drug of choice.<br />

Metformin blocks hepatic gluconeogenesis by an unknown mechanism, but has<br />

minimal effect on insulin resistance. It also reduces the prothrombotic state by<br />

an effect on plasminogen activator inhibitor type 1 (PAI-1), and its early use can<br />

prevent the progression from prediabetes to diabetes. Metformin’s effect on CVD<br />

risk is uncertain, although it does cause a modest weight loss.


Obesity, Nutrigenomics, Metabolic Syndrome, and Type 2 Diabetes 115<br />

Drugs aimed at increasing insulin secretion (e.g., the sulfonylureas) have<br />

shown no significant impacts on CVD risk and may even increase risk after a<br />

myocardial infarction (MI).<br />

Insulin limits markers of inflammation, reduces intimal medial thickness<br />

(IMT, a surrogate marker of CVD risk), and has been shown to lower risk after<br />

coronary artery bypass and to reduce post-MI mortality. On the whole, intensive<br />

treatment of hyperglycemia per se has been seen to have minimal effect on CVD<br />

risk. 15<br />

Besides insulin, metformin, and sulfonylureas, a new class of drug has<br />

emerged in the last decade, the thiazolidinediones (TZDs) or glitazones. These<br />

drugs are uniquely specific in reversing insulin resistance as agonists of the<br />

peroxisome proliferator-activated receptor gamma (PPAR-gamma). This is a<br />

nuclear receptor found in adipocytes and elsewhere that regulates the expression<br />

of genes involved in lipid and glucose metabolism, vascular function, thrombotic<br />

control, and inflammation. However, glitazones can also stimulate adipogenesis<br />

and so are used as second-line agents after metformin which encourages weight<br />

loss.<br />

Exenatide is a new injectable treatment developed after observation of the<br />

“incretin effect” in which insulin levels were seen to rise significantly more after<br />

oral glucose than after an IV injection of glucose due to the action of a thenunidentified<br />

gut peptide. 16 Exenatide is a synthetic version of a salivary peptide<br />

in the Gila monster, and it mimics an endogenous incretin, glucagon-like peptide<br />

1 in stimulating glucose-dependent insulin secretion, regulating gastric emptying,<br />

and inhibiting glucagon secretion, food intake, and acute plasma glucose.<br />

Exenatide is indicated as an adjunct to metformin and has been associated with<br />

weight loss as well.<br />

Another new injectable treatment is pramlintide, a more soluble analog of<br />

amylin, which is co-secreted by the pancreatic gamma cell with insulin. Amylin,<br />

and thus pramlintide, promotes satiety, inhibits glucagon release, and delays<br />

gastric emptying, all through an effect via the central nervous system. 16<br />

BETA CELL FAILURE <strong>AND</strong> TYPE 2 DIABETES MELLITUS<br />

Type 2 diabetes involves a progressive defect of insulin secretion that precedes<br />

the development of hyperglycemia. 17 This defect appears to be at least in part due<br />

to a deficit in beta cell mass. 18–20 Several therapeutic strategies now being proposed<br />

may reverse the defect in beta cell mass in people with type 2 diabetes, for<br />

example, glucagon-like peptide 1 or glucagon-like peptide 1–like surrogates. 21<br />

In humans there is a curvilinear relationship between the relative beta cell<br />

volume (and presumably the beta cell mass) and fasting blood glucose concentration.<br />

The present data reveal a narrow range of blood glucose over a wide range<br />

of fractional beta cell volume (up to ~10%) and then a much wider range of blood<br />

glucose values over a narrow range of volumes at low beta cell volumes, with<br />

the threshold set by the curve at ~1.1% defining that difference. These findings<br />

imply a much greater tolerance for variance in insulin sensitivity above this


116 Oxidative Stress and Inflammatory Mechanisms<br />

threshold and that below-the-threshold variance in insulin sensitivity and functional<br />

defects in insulin secretion have a much greater impact on blood glucose.<br />

Autopsy studies involve some important limitations. The numbers of cases<br />

are frequently relatively small. The studies are inevitably cross-sectional and<br />

retrospective. Butler and others have presented data suggesting that the decline<br />

in beta cell mass in type 2 diabetes is caused by increased beta cell apoptosis. 19,22<br />

Therefore, these data suggest that inhibition of beta cell apoptosis to avoid a beta<br />

cell deficit may be effective to delay and/or avoid the onset of diabetes. Indeed,<br />

both metformin and TZDs have been reported to inhibit beta cell apoptosis in<br />

vitro 22,23 and delay onset of type 2 diabetes in clinical studies. 24,25<br />

The potential mechanisms underlying increased beta cell apoptosis in type 2<br />

diabetes include toxicity from islet amyloid polypeptide oligomer formation, free<br />

fatty acids (lipotoxicity), free oxygen radical toxicity, and, once hyperglycemia<br />

supervenes, glucose-induced apoptosis (glucotoxicity). 26,27 The steep increase in<br />

blood glucose concentration with beta cell deficiency is consistent with the well<br />

known deleterious effects of hyperglycemia per se on beta cell function. These<br />

include defective glucose sensing due to reduced glucokinase activity, 28 impaired<br />

glucose-induced insulin secretion due to increased uncoupling protein 2 activity, 29<br />

and depletion of immediately secretable insulin stores. 30 Also, hyperglycemia<br />

reduces insulin sensitivity, further compounding the effects of decreased insulin<br />

secretion. 31 While these observations suggest that relatively small increases in<br />

beta cell mass may have useful actions in restoring blood glucose control, it is<br />

likely that aggressive normalization of blood glucose concentrations is required<br />

to accompany any strategy to increase beta cell mass to overcome the deleterious<br />

effects of hyperglycemia as well as glucose-induced beta cell apoptosis.<br />

ADIPOKINES<br />

Adipose tissue secretes bioactive peptides, termed adipokines, which act locally<br />

and distally through autocrine, paracrine and endocrine effects. In obesity,<br />

increased production of most adipokines impacts on multiple functions such as<br />

appetite and energy balance, immunity, insulin sensitivity, angiogenesis, blood<br />

pressure, lipid metabolism and hemostasis, all of which are linked with cardiovascular<br />

disease. Enhanced activities of TNF and IL-6 are involved in the development<br />

of obesity-related insulin resistance. Angiotensinogen has been implicated<br />

in hypertension and PAI-1 in impaired fibrinolysis.<br />

Other adipokines like adiponectin and leptin, at least in physiological concentrations,<br />

are insulin sparing as they stimulate beta oxidation of fatty acids in<br />

skeletal muscle. The role of resistin is less understood. It is implicated in insulin<br />

resistance in rats, but probably not in humans. Reducing adipose tissue mass<br />

through weight loss in association with exercise can lower TNF-α and IL-6 levels<br />

and increase adiponectin concentrations, whereas drugs such as TZDs increase<br />

endogenous adiponectin production. In-depth understanding of the pathophysiology<br />

and molecular actions of adipokines may in the coming years lead to


Obesity, Nutrigenomics, Metabolic Syndrome, and Type 2 Diabetes 117<br />

effective therapeutic strategies designed to protect against atherosclerosis in obese<br />

patients.<br />

Obesity associated with unfavorable changes in adipokine expression such<br />

as increased levels of TNF-α, IL-6, resistin, PAI-1 and leptin, and reduced levels<br />

of adiponectin affects glycemic homeostasis, vascular endothelial function, and<br />

the coagulation system, thus accelerating atherosclerosis. Adipokines and a lowgrade<br />

inflammatory state may be the link between the metabolic syndrome with<br />

its cluster of obesity and insulin resistance and cardiovascular disease.<br />

In fact, atherosclerosis is now recognized as an inflammatory process of the<br />

arterial wall. Monocytes adhere to the endothelium and then migrate into the<br />

subendothelial space where they become foam cells loaded with oxidized lipoproteins.<br />

Foam cell production of metalloproteinases leads to rupture of the<br />

atherosclerotic plaque’s fibrous cap and then to rupture of the plaque itself. 32<br />

Thus, an inflammatory process accounts for both the development and evolution<br />

of atherosclerosis.<br />

In this inflammatory process, adipokines play multiple roles. TNF-α activates<br />

the transcription factor nuclear factor-κβ, with subsequent inflammatory changes<br />

in vascular tissue. These include increased expression of intracellular adhesion<br />

molecule (ICAM)-1 and vascular cell adhesion molecule (VCAM)-1, 33,34 which<br />

enhances monocyte adhesion to the vessel wall, greater production of MCP-1<br />

and M-CSF from endothelial cells and vascular smooth muscle cells, 35,36 and upregulated<br />

macrophage expression of inducible nitric oxide (NO) synthase, interleukins,<br />

superoxide dismutase, etc. 37,38 Leptin, especially in the presence of high<br />

glucose, stimulates macrophages to accumulate cholesterol. 39 IL-6 exerts proinflammatory<br />

activity in itself and by increasing IL-1 and TNF-α. 40 Importantly,<br />

IL-6 also stimulates liver production of C-reactive protein which is considered a<br />

predictor of atherosclerosis. 41 IL-6 may also influence glucose tolerance by regulation<br />

of visfatin. Visfatin, a newly discovered adipocytokine in human visceral<br />

fat, exerts insulin-mimetic effects in cultured cells and lowers plasma glucose<br />

levels in mice through activation of the insulin receptor. 42<br />

PAI-1 concentrations regulated by transcription factor nuclear factor-κβ are<br />

abnormally high in hyperglycemia, obesity, and hypertriglyceridemia, 43 because<br />

of increased PAI-1 gene expression. 44 PAI-1 inhibits fibrin clot breakdown,<br />

thereby favoring thrombus formation upon ruptured atherosclerotic plaques. 45 In<br />

humans, circulating PAI-1 levels correlate with atherosclerotic events and mortality,<br />

and some studies suggest PAI-1 is an independent risk factor for coronary<br />

artery disease. 46 Angiotensinogen is a precursor of angiotensin II (AngII), which<br />

stimulates ICAM-1, VCAM-1, MCP-1, and M-CSF expression in vessel wall<br />

cells. 47 AngII also reduces NO bioavailability 48 with loss of vasodilator capacity<br />

and increased platelet adhesion to vessel walls.<br />

In humans, endothelial dysfunction is indicative of the preclinical stages of<br />

atherosclerosis and is prognostic of future cardiovascular events. 49,50 High concentrations<br />

of pro-inflammatory adipokines may contribute to development of<br />

endothelial dysfunction. At this stage of disease, the role of resistin is particularly<br />

interesting. In vitro studies show resistin activates endothelial cells which, when


118 Oxidative Stress and Inflammatory Mechanisms<br />

incubated with recombinant human resistin, release more endothelin-1 and<br />

VCAM-1. 51 Recombinant human resistin is also reported to induce higher expression<br />

of mRNA of VCAM, ICAM-1, and pentraxin-3 from endothelial cells, 52 thus<br />

expressing a biochemical pattern of dysfunctional endothelium. Finally, resistin<br />

also induces proliferation of aortic smooth muscle cells. 53 In asymptomatic<br />

patients with family histories of coronary heart disease, plasma resistin levels are<br />

predictive of coronary atherosclerosis even after control for other established risk<br />

factors. 54,55<br />

CONCLUSION<br />

The molecular effects of adipokines represent a challenging area of research and<br />

in-depth understanding of their pathophysiology and molecular actions will<br />

undoubtedly lead to the discovery of effective therapeutic interventions. Reducing<br />

adipose tissue mass and consequently adipokine concentrations will prevent the<br />

metabolic syndrome and, if the hypothesis of adipokine-related linkage with<br />

atherosclerosis is proven, help prevent the development of atherosclerosis. Despite<br />

the new findings in the field of adipokines, researchers are still led to focus back<br />

on obesity as an essential primary target in the continued effort to reduce the risk<br />

of developing the metabolic syndrome and type 2 diabetes with its associated<br />

cardiovascular complications.<br />

REFERENCES<br />

1. Wilson PWF et al. Overweight and obesity as determinants of cardiovascular risk:<br />

the Framingham experience. Arch Intern Med 2002; 162: 1867.<br />

2. Centers for Disease Control and Prevention, Department of Health and Human<br />

Services. Overweight and obesity: U.S. obesity trends, 1985–2004. Available at:<br />

http://www.cdc.gov/nccdphp/dnpa/obesity/trend/maps/. Accessed April 13, 2006.<br />

3. National Heart, Lung, and Blood Institute, North American Association for the<br />

Study of Obesity. The practical guide: identification, evaluation, and treatment of<br />

overweight and obesity in adults. Available at http://www.nhlbi.nih.gov/guidelines/obesity/practgde.htm.<br />

Accessed April 13, 2006.<br />

4. Shick SM et al. Persons successful at long-term weight loss and maintenance<br />

continue to consume a low calorie, low fat diet. J Am Dietetic Assoc 1998; 98: 408.<br />

5. Dansinger ML et al. Comparison of the Atkins, Ornish, Weight Watchers, and<br />

Zone diets for weight loss and heart disease risk reduction: a randomized trial.<br />

JAMA 2005; 293: 43.<br />

6. Hultquist CN, Albright C, and Thompson DL. Comparison of walking recommendations<br />

in previously inactive women. Med Sci Sports Exerc 2005; 37: 676.<br />

7. Thomas RA and Dlugosz CK. Tools and techniques for fighting the obesity<br />

epidemic. Program and abstracts of the American Pharmacists Association Annual<br />

Meeting and Exposition; March 17–21, 2006, San Francisco.<br />

8. Grundy S. Metabolic syndrome: connecting and reconciling cardiovascular and<br />

diabetes worlds. J Am Coll Cardiol 2006; 47: 1093. Available at Medscape Cardiology,<br />

http://www.medscape.com/viewarticle/524081. Accessed April 13, 2006.


Obesity, Nutrigenomics, Metabolic Syndrome, and Type 2 Diabetes 119<br />

9. Grundy SM et al. Diagnosis and management of metabolic syndrome: an American<br />

Heart Association/National Heart, Lung, and Blood Institute scientific statement.<br />

Circulation 2005; 112: e285.<br />

10. Expert Panel on Detection, Evaluation, and Treatment of High Blood Cholesterol<br />

in Adults. Executive Summary of the Third Report of National Cholesterol Education<br />

Program (NCEP) Expert Panel on Detection, Evaluation, and Treatment of<br />

High Blood Cholesterol in Adults (Adult Treatment Panel III). JAMA 2001; 285:<br />

2486.<br />

11. Kendall DM and Harmel AP. The metabolic syndrome, type 2 diabetes, and<br />

cardiovascular disease: understanding the role of insulin resistance. Am J Manag<br />

Care 2002; 8: S635.<br />

12. Curran JE et al. Genetic variation in selenoprotein S influences inflammatory<br />

response. Nat Genet 2005; 37: 1234.<br />

13. Rankinen T et al. The human obesity gene map: the 2005 update. Obesity 2006;<br />

14: 529.<br />

14. Schulz LO et al. Effects of traditional and Western environments on prevalence<br />

of type 2 diabetes in Pima indians in Mexico and the U.S. Diabetes Care 2006;<br />

29: 1866.<br />

15. United Kingdom Prospective Diabetes Study 24: a 6-year, randomized, controlled<br />

trial comparing sulfonylurea, insulin, and metformin therapy in patients with<br />

newly diagnosed type 2 diabetes that could not be controlled with diet therapy.<br />

Ann Intern Med 1998; 128: 165.<br />

16. Aronoff SL et al. Glucose metabolism and regulation: beyond insulin and glucagons.<br />

Diabetes Spectrum 2004; 17: 183.<br />

17. United Kingdom Prospective Diabetes Study Group. Study 16: overview of 6<br />

years’ therapy of type II diabetes: a progressive disease. Diabetes 1995; 44: 1249.<br />

18. Kloppel G et al. Islet pathology and the pathogenesis of type 1 and type 2 diabetes<br />

mellitus revisited. Surv Synth Pathol Res 1985; 4: 110.<br />

19. Butler AE et al. Beta cell deficit and increased beta cell apoptosis in humans with<br />

type 2 diabetes. Diabetes 2003; 52: 102.<br />

20. Yoon KH et al. Selective beta-cell loss and alpha-cell expansion in patients with<br />

type 2 diabetes mellitus in Korea. J Clin Endocrinol Metab 2003; 88: 2300.<br />

21. Nauck MA. Glucagon-like peptide 1 (GLP-1) in the treatment of diabetes. Horm<br />

Metab Res 2004; 36: 852.<br />

22. Marchetti P et al. Pancreatic islets from type 2 diabetic patients have functional<br />

defects and increased apoptosis that are ameliorated by metformin. J Clin Endocrinol<br />

Metab 2004; 89: 5535.<br />

23. Zeender E et al. Pioglitazone and sodium salicylate protect human beta-cells<br />

against apoptosis and impaired function induced by glucose and interleukin-1beta.<br />

J Clin Endocrinol Metab 2004; 89: 5059.<br />

24. Knowler WC et al. Reduction in the incidence of type 2 diabetes with lifestyle<br />

intervention or metformin. New Engl J Med 2002; 346: 393.<br />

25. Buchanan TA et al. Preservation of pancreatic ß-cell function and prevention of<br />

type 2 diabetes by pharmacological treatment of insulin resistance in high-risk<br />

hispanic women. Diabetes 2002; 51: 2796.<br />

26. Butler AE et al. Diabetes due to a progressive defect in ß-cell mass in rats<br />

transgenic for human islet amyloid polypeptide (HIP Rat): a new model for type<br />

2 diabetes. Diabetes 2004; 53: 1509.


120 Oxidative Stress and Inflammatory Mechanisms<br />

27. Donath MY and Halban PA. Decreased beta-cell mass in diabetes: significance,<br />

mechanisms and therapeutic implications (review). Diabetologia 2004; 47: 581.<br />

28. Matschinsky FM, Glaser B, and Magnuson MA. Pancreatic ß-cell glucokinase:<br />

closing the gap between theoretical concepts and experimental realities. Diabetes<br />

1998; 47: 307.<br />

29. Krauss S et al. Superoxide-mediated activation of uncoupling protein 2 causes<br />

pancreatic beta cell dysfunction. J Clin Invest 2003; 112: 1831.<br />

30. Ritzel RA et al. Induction of beta-cell rest by a Kir6.2/SUR1-selective K(ATP)channel<br />

opener preserves beta-cell insulin stores and insulin secretion in human<br />

islets cultured at high (11 mM) glucose. J Clin Endocrinol Metab 2004; 89: 795.<br />

31. Tomas E et al. Hyperglycemia and insulin resistance: possible mechanisms. Ann<br />

NY Acad Sci 2002; 967: 43.<br />

32. <strong>Lib</strong>by, P. Changing concepts of atherogenesis. J Int Med 2000; 247, 349.<br />

33. Landry DB et al. Activation of the NF- B and I-B system in smooth muscle cells<br />

after rat arterial injury. Induction of vascular cell adhesion molecule-1 and monocyte<br />

chemoattractant protein-1. Am J Pathol 1997; 151, 1085.<br />

34. Iademarco MF, McQuillan JJ, and Dean DC. Vascular cell adhesion molecule 1:<br />

contrasting transcriptional control mechanisms in muscle and endothelium. Proc<br />

Natl Acad Sci USA 1993; 90, 3943.<br />

35. Eck SL et al. Inhibition of phorbol ester-induced cellular adhesion by competitive<br />

binding of NF-B in vivo. Mol Cell Biol 1993; 13, 6530.<br />

36. Clesham GJ et al. High adenoviral loads stimulate NFB-dependent gene expression<br />

in human vascular smooth muscle cells. Gene Ther 1998; 174.<br />

37. Xie QW, Kashiwabara Y, and Nathan C. Role of transcription factor NF-B/Rel in<br />

induction of nitric oxide synthase. J Biol Chem 1994; 269, 4705.<br />

38. Goto M et al. Involvement of NF-B p50/p65 heterodimer in activation of the<br />

human prointerleukin-1 gene at two subregions of the upstream enhancer element.<br />

Cytokine 1999; 11, 16.<br />

39. O'Rourke L et al. Glucose-dependent regulation of cholesterol ester metabolism<br />

in macrophages by insulin and leptin. J Biol Chem 2002; 277, 42557.<br />

40. Yudkin JS et al. Inflammation, obesity, stress and coronary heart disease: is<br />

interleukin-6 the link? Atherosclerosis 2000; 148, 209.<br />

41. Ridker PM et al. Comparison of C-reactive protein and low-density lipoprotein<br />

cholesterol levels in the prediction of first cardiovascular events. New Engl J Med<br />

2002; 347, 1557.<br />

42. Fukuhara A et al. Visfatin: a protein secreted by visceral fat that mimics the effects<br />

of insulin. Science 2005; 307, 426.<br />

43. Stentz FB et al. Proinflammatory cytokines, markers of cardiovascular risks,<br />

oxidative stress, and lipid peroxidation in patients with hyperglycemic crises.<br />

Diabetes 2004; 53, 2079.<br />

44. Gabriely I et al. Hyperglycemia induces PAI-1 gene expression in adipose tissue<br />

by activation of the hexosamine biosynthetic pathway. Atherosclerosis 2002; 160,<br />

115.<br />

45. Sobel BE. Increased plasminogen activator inhibitor-1 and vasculopathy: a reconcilable<br />

paradox. Circulation 1999; 99, 2496.<br />

46. Thogersen AM et al. High plasminogen activator inhibitor and tissue plasminogen<br />

activator levels in plasma precede a first acute myocardial infarction in both men<br />

and women: evidence for the fibrinolytic system as an independent primary risk<br />

factor. Circulation 1998; 98, 2241.


Obesity, Nutrigenomics, Metabolic Syndrome, and Type 2 Diabetes 121<br />

47. Tham DM et al. Angiotensin II is associated with activation of NF-B-mediated<br />

genes and downregulation of PPARs. Physiol Genomics 2002; 11, 21.<br />

48. Cai H et al. NAD (P) H oxidase-derived hydrogen peroxide mediates endothelial<br />

nitric oxide production in response to angiotensin II. J Biol Chem 2002; 277,<br />

48311.<br />

49. Widlansky ME et al. The clinical implications of endothelial dysfunction. J Am<br />

Coll Cardiol 2003; 42, 1149.<br />

50. Jambrik Z et al. Peripheral vascular endothelial function testing for the diagnosis<br />

of coronary artery disease. Am Heart J 2004; 148, 684.<br />

51. Verma S et al. Resistin promotes endothelial cell activation: further evidence of<br />

adipokine–endothelial interaction. Circulation 2003; 108, 736.<br />

52. Kawanami D et al. Direct reciprocal effects of resistin and adiponectin on vascular<br />

endothelial cells: a new insight into adipocytokine–endothelial cell interactions.<br />

Biochem Biophys Res Commun 2004; 314, 415.<br />

53. Calabro P et al. Resistin promotes smooth muscle cell proliferation through<br />

activation of extracellular signal-regulated kinase 1/2 and phosphatidylinositol 3kinase<br />

pathways. Circulation 2004; 110, 3335.<br />

54. Pinkney JH et al. Endothelial dysfunction: cause of the insulin resistance syndrome.<br />

Diabetes 1997; 46, S9.<br />

55. Reilly MP et al. Resistin is an inflammatory marker of atherosclerosis in humans.<br />

Circulation 2005; 111, 932.


8<br />

CONTENTS<br />

Post-Prandial Endothelial<br />

Dysfunction, Oxidative<br />

Stress, and Inflammation<br />

in Type 2 Diabetes<br />

Antonio Ceriello<br />

Abstract .............................................................................................................123<br />

Introduction .......................................................................................................124<br />

Possible Role of Hyperglycemic Spikes in Cardiovascular Diseases..............124<br />

Fasting Hyperglycemia and Cardiovascular Disease..............................124<br />

Post-Prandial Hyperglycemia and Cardiovascular Disease:<br />

Epidemiological Evidence...........................................................125<br />

Post-Prandial Hyperglycemia and Cardiovascular Disease:<br />

Intervention Studies.....................................................................126<br />

Mechanisms Involved........................................................................................127<br />

Post-Prandial Hyperglycemia and Oxidative and Nitrosative Stress...............128<br />

Conclusions .......................................................................................................131<br />

References .........................................................................................................132<br />

ABSTRACT<br />

Increasing evidence suggests that the post-prandial state is a contributing factor to<br />

the development of atherosclerosis. In diabetes, the post-prandial phase is characterized<br />

by a rapid and large increase in blood glucose levels. The possibility that<br />

the post-prandial “hyperglycemic spikes” may be relevant to the onset of cardiovascular<br />

complications recently received much attention. Epidemiological studies<br />

and preliminary intervention studies have shown that post-prandial hyperglycemia<br />

is a direct and independent risk factor for cardiovascular disease. Most of the<br />

cardiovascular risk factors are modified in the post-prandial phase in diabetic subjects<br />

and directly affected by an acute increase of glycemia. The mechanisms<br />

through which acute hyperglycemia exerts its effects may be identified in the<br />

123


124 Oxidative Stress and Inflammatory Mechanisms<br />

production of free radicals. This alarmingly suggestive body of evidence for a<br />

harmful effect of post-prandial hyperglycemia on diabetic complications has been<br />

sufficient to influence guidelines from key professional scientific societies. Correcting<br />

post-prandial hyperglycemia may form part of a strategy for the prevention and<br />

management of cardiovascular diseases in diabetes.<br />

INTRODUCTION<br />

Diabetes mellitus is characterized by a high incidence of cardiovascular disease 1<br />

and poor control of hyperglycemia appears to play a significant role in the<br />

development of cardiovascular disease in diabetes. 2 Increasing evidence indicates<br />

that the post-prandial state is an important contributing factor to the development<br />

of atherosclerosis. 3 In diabetes, the post-prandial phase is characterized by a rapid<br />

and large increase in blood glucose levels. The possibility that these post-prandial<br />

“hyperglycemic spikes” may be relevant to the pathophysiology of late diabetic<br />

complications is recently receiving much attention.<br />

In this chapter, epidemiological data and preliminary results of intervention<br />

studies indicating that post-prandial hyperglycemia represents an increased risk<br />

for cardiovascular disease are surveyed. The proposed mechanisms involved in<br />

this effect are summarized.<br />

POSSIBLE ROLE OF HYPERGLYCEMIC SPIKES IN<br />

CARDIOVASCULAR <strong>DISEASE</strong>S<br />

FASTING HYPERGLYCEMIA <strong>AND</strong> CARDIOVASCULAR <strong>DISEASE</strong><br />

Over the last 10 years, many studies showed an independent relationship between<br />

cardiovascular diseases and glycemic control in patients with type 2 diabetes. 2<br />

These studies involved thousands of subjects, often newly diagnosed, who were<br />

followed up for periods ranging from 3.5 to 11 years and were evaluated on the<br />

basis of various cardiovascular end-points. 2 It is necessary to underline that the<br />

majority of these studies used a single baseline fasting glycemic value or a single<br />

value of glycated hemoglobin A1c (HbA1c) to predict cardiovascular events<br />

occurring many years later.<br />

For instance, the observational version of the United Kingdom Prospective<br />

Diabetes Study (UKPDS) showed that the mean HbA1c value was a good predictor<br />

of ischemic heart disease. 4 In particular, multivariate analysis showed that<br />

for each 1% increment in HbA1c there was an approximate 10% increase in the<br />

risk of coronary heart disease. 4 This evidence is not substantially different compared<br />

with the results of the interventional version of the UKPDS. In this trial,<br />

even though the result was not significant (p


Dysfunction, Stress, and Inflammation in Type 2 Diabetes 125<br />

improving insulin resistance, may have significantly improved the “cluster” of<br />

cardiovascular risk factors associated with insulin resistance.<br />

The relationship existing between macroangiopathy and fasting plasma glucose<br />

or HbA1c is weaker than that observed with microangiopathy. 2 This was<br />

found in cross-sectional or longitudinal studies. These data support the hypothesis<br />

that fasting plasma glucose or HbA1c alone is unable to describe thoroughly<br />

the glycemic disorders occurring in diabetes and its impact on cardiovascular<br />

disease. In addition to fasting glycemia and HbA1c, emphasis has recently been<br />

given to the relationship between post-prandial hyperglycemia and cardiovascular<br />

diseases.<br />

POST-PR<strong>AND</strong>IAL HYPERGLYCEMIA <strong>AND</strong> CARDIOVASCULAR <strong>DISEASE</strong>:<br />

EPIDEMIOLOGICAL EVIDENCE<br />

The oral glucose tolerance test (OGTT) has been used mostly in epidemiological<br />

studies that attempt to evaluate the risk of cardiovascular disease. The main<br />

advantage of the OGTT is its simplicity: a single plasma glucose measurement<br />

2 hours after a glucose load determines whether glucose tolerance is normal,<br />

impaired, or indicates overt diabetes.<br />

The caveats of the OGTT are numerous because 75 or 100 g glucose is almost<br />

never ingested during a meal and, more importantly, many events associated with<br />

ingesting a pure glucose solution do not incorporate the numerous metabolic<br />

events associated with eating a mixed meal. Moreover, the relationship between<br />

glycemia and meal content is contingent upon the contents of the meal. 7 However,<br />

it has recently been demonstrated that the level of glycemia reached 2 hours after<br />

an OGTT was closely related to the level of glycemia after a standardized meal<br />

(mixed meal in the form of wafers containing oat fractionation products, soy<br />

protein, and canola oil sweetened with honey: 345 kcal, 10.7 g fat, 12.1 g protein,<br />

8.9 g simple sugars, 41.1 g starch, and 3.8 g dietary fiber), suggesting that the<br />

OGTT may represent a valid tool to reveal altered carbohydrate metabolism<br />

during a meal. 8 Interestingly, the correlation is more consistent for the values of<br />

glycemia in the impaired glucose tolerance range. 8<br />

From an epidemiological point of view, the Hoorn Study, 9 the Honolulu Heart<br />

Study, 10 the Chicago Heart Study, 11 and the more recent DECODE Study 12 have<br />

clearly shown that glucose serum levels 2 hours after oral challenge with glucose<br />

are powerful predictors of cardiovascular risk. This evidence is confirmed also<br />

by two important meta-analyses: the first by Coutinho et al. examined studies of<br />

95,783 subjects. 13 The second, covering more than 20,000 subjects, pooled the<br />

data of the Whitehall Study, the Paris Prospective Study, and the Helsinki Policemen<br />

Study. 14<br />

The possible role of post-prandial hyperglycemia as an independent risk factor<br />

has also been supported by the Diabetes Intervention Study showing how in type<br />

2 diabetics post-prandial hyperglycemia predicts infarction 15 and by another<br />

study associating post-prandial hyperglycemia levels with medio-intimal carotid<br />

thickening. 16


126 Oxidative Stress and Inflammatory Mechanisms<br />

TABLE 8.1<br />

Epidemiological Studies Showing Association of Post-Prandial<br />

Hyperglycemia with Risk of CVD and Mortality<br />

Study Result<br />

Intriguing evidence comes from a study that demonstrates how medio-intimal<br />

carotid thickening is correlated not only with post-prandial glucose serum level<br />

but particularly with glycemic spikes during the OGTT. 17 A post-challenge glucose<br />

spike was defined as the difference between the maximal post-challenge<br />

glucose level during OGTT, irrespective of the time after glucose challenge, and<br />

the level of fasting plasma glucose. 17 Epidemiological studies are summarized in<br />

Table 8.1.<br />

Indirect evidence of the unfavorable role of acute hyperglycemia on cardiovascular<br />

diseases is also available. Hyperglycemia during a cardiovascular acute<br />

event is unfavorable from a prognostic view in cases both of myocardial<br />

infarction 18,19 and stroke. 20,21 A worst prognosis has been demonstrated for both<br />

cases in diabetic and non-diabetic subjects. 18–21<br />

Regarding infarction, it has been recently demonstrated by a meta-analysis<br />

that there is a continuous correlation between glucose serum levels and the<br />

seriousness of prognosis even in non-diabetic subjects, 22 while intensive insulin<br />

treatment during acute myocardial infarction reduced long-term mortality in<br />

diabetic patients. 23 This is consistent with the evidence that an acute increase of<br />

glycemia significantly prolongs the QT in normal subjects 24 and that during<br />

myocardial infarction increased glucose level is capable of inducing such electrophysiological<br />

alterations as to favor the occurrence of arrhythmias whose<br />

outcomes could even be fatal. 25<br />

POST-PR<strong>AND</strong>IAL HYPERGLYCEMIA <strong>AND</strong> CARDIOVASCULAR <strong>DISEASE</strong>:<br />

INTERVENTION STUDIES<br />

Ref.<br />

No.<br />

Hoorn Study 2-hour glucose better predictor of mortality than HbA1c 9<br />

Honolulu Heart Program 1-hour glucose predicts coronary heart disease 10<br />

Chicago Heart Study 2-hour post-challenge glucose predicts all cause mortality 11<br />

DECODE High 2-hour post-load blood glucose associated with<br />

increased risk of death independent of fasting glucose<br />

12<br />

Coutinho M et al. 2-hour glucose associated with CHD 13<br />

Whitehall, Paris, and 2-hour post-challenge glucose predicts all cause and CHD 14<br />

Helsinki Studies<br />

mortality<br />

Diabetes Intervention Study Post-meal, but not fasting glucose, associated with CHD 15<br />

One of the major concerns about the role of post-prandial hyperglycemia in<br />

cardiovascular disease has been, until now, the absence of intervention studies.<br />

That situation is changing.


Dysfunction, Stress, and Inflammation in Type 2 Diabetes 127<br />

The STOP-NIDDM trial has presented data indicating that treatment of<br />

patients who have IGT with acarbose, an α-glucosidase inhibitor that specifically<br />

reduces post-prandial hyperglycemia, is associated not only with a 36% reduction<br />

in the risk of progression to diabetes 26 but also with a 34% risk reduction in the<br />

development of new cases of hypertension and a 49% risk reduction in cardiovascular<br />

events. 27 In addition, in a subgroup of patients, carotid intima media<br />

thickness (CIMT) was measured before randomization and at the end of the<br />

study. 28 Acarbose treatment was associated with a significant decrease in the<br />

progression of intima media thickness, an accepted surrogate for atherosclerosis. 28<br />

In a recent meta-analysis of type 2 diabetic patients, acarbose treatment was<br />

associated with significant reductions in cardiovascular events, even after adjusting<br />

for other risk factors. 29 Finally, very recently, the effects of two insulin<br />

secretagogues, repaglinide and glyburide, known to have different efficacies for<br />

post-prandial hyperglycemia, CIMT, and markers of systemic vascular inflammation<br />

in type 2 diabetic patients, have been evaluated. 30 After 12 months,<br />

post-prandial glucose peak measurements were 148 ± 28 mg/dL in the repaglinide<br />

group and 180 ± 32 mg/dL in the glyburide group (p 0.020 mm, was observed in 52% of diabetics receiving repaglinide and in<br />

18% of those receiving glyburide (p


128 Oxidative Stress and Inflammatory Mechanisms<br />

2 diabetic patients and that the decrease correlated inversely with the magnitude<br />

of post-prandial hyperglycemia. 39<br />

The possible role of hyperglycemia in the activation of blood coagulation has<br />

previously been reviewed. 40 It emerges that acute glycemic variations are matched<br />

with a series of alterations of coagulation that are likely to cause a thrombosis.<br />

This tendency is documented by studies that demonstrate how inducing hyperglycemia<br />

produces a shortening of fibrinogen half-life 41 and increases in fibrinopeptide<br />

A, 42,43 fragments of prothrombin, 44 factor VII, 45 and platelet aggregation 46<br />

in both normal and diabetic subjects. These data indicate that coagulation is<br />

activated during experimental hyperglycemia. It is interesting that the over-production<br />

of thrombin by post-prandial hyperglycemia in diabetics has already been<br />

already documented. The phenomenon is strictly dependent on the glycemic levels<br />

reached. 47<br />

Adhesion molecules regulate the interactions of endothelium and leukocytes.<br />

48 They participate in the process of atherogenesis because their greater<br />

expression would imply an increase in the adhesion of leukocytes (monocytes in<br />

particular) to the endothelium. 49 It is well known that this is considered one of<br />

the early stages of the process leading to atheromatous lesions. Among the various<br />

pro-adhesive molecules, ICAM-1 has received particular interest. Increases in the<br />

circulating form of this molecule have been demonstrated in subjects with vascular<br />

disease 50 and with diabetes mellitus with or without vascular disease. 51,52<br />

These increases have been considered indications of the activation of the atherogenic<br />

process.<br />

The soluble form of ICAM-1 is stored in cells and can be quickly expressed<br />

outside them as a consequence of various stimuli. It has been demonstrated that<br />

acute hyperglycemia in both normal and diabetic subjects is a sufficient stimulus<br />

for the circulating level of ICAM-1 to increase, thus activating one of the first<br />

stages of the atherogenic process. 53,54<br />

The concept of atherosclerosis as an inflammatory disease even in diabetes<br />

is now well established. 55 Studies support the evidence that acute hyperglycemia<br />

during a hyperglycemic clamp 56 or in a post-prandial state 57 can increase the<br />

production of plasma IL-6, TNF-α, and IL-18.<br />

POST-PR<strong>AND</strong>IAL HYPERGLYCEMIA <strong>AND</strong> <strong>OXIDATIVE</strong> <strong>AND</strong><br />

NITROSATIVE <strong>STRESS</strong><br />

Recent studies demonstrated that hyperglycemia induces an over-production of<br />

superoxide by the mitochondrial electron-transport chain. 58 Superoxide overproduction<br />

is accompanied by increased nitric oxide generation due to eNOS and<br />

iNOS in an uncoupled state — a phenomenon favoring the formation of peroxynitrite,<br />

a strong oxidant that in turn damages DNA. 59 DNA damage is an obligatory<br />

stimulus for the activation of the poly(ADP-ribose) polymerase nuclear<br />

enzyme. Poly(ADP-ribose) polymerase activation in turn depletes the intracellular<br />

concentration of its substrate NAD + , slowing the rate of glycolysis, electron


Dysfunction, Stress, and Inflammation in Type 2 Diabetes 129<br />

Hyperglycemia<br />

Mitochondria<br />

− O2 PKC NF-kB<br />

NAD(P)H<br />

oxidase<br />

−<br />

O2 PARP<br />

NAD +<br />

DNA damage<br />

iNOS eNOS<br />

NO<br />

GAPDH<br />

Peroxynitrite<br />

Nitrotyrosine<br />

Endothelial dysfunction<br />

Polyol pathway<br />

AGE formation<br />

Hexosamine Flux<br />

Adhesion molecules<br />

Proinflammatory<br />

Cytokines<br />

Diabetic<br />

complications<br />

FIGURE 8.1 In endothelial cells, glucose can pass freely in an insulin-independent manner<br />

through the cell membrane. Intracellular hyperglycemia induces over-production of<br />

superoxide at the mitochondrial level. Over-production of superoxide is the first and key<br />

event in the activation of all other pathways involved in the pathogenesis of diabetic<br />

complications such as polyol pathway flux, increased AGE formation, activation of protein<br />

kinase C and NF-kB, and increased hexosamine pathway flux. O 2 – reacting with NO<br />

produces peroxynitrite (ONOO – ). Superoxide over-production reduces eNOS activity but<br />

through NF-kB and PKC activates NAD(P)H and increases iNOS expression: the final<br />

effect is an increased nitric oxide generation. This condition favors the formation of the<br />

strong peroxynitrite oxidant that in turn produces in iNOS and eNOS an uncoupled state,<br />

resulting in the production of superoxide rather than NO, and damages DNA. DNA damage<br />

is an obligatory stimulus for the activation of the poly(ADP-ribose) polymerase nuclear<br />

enzyme. Poly(ADP-ribose) polymerase activation in turn depletes the intracellular concentration<br />

of its substrate NAD + , slowing the rate of glycolysis, electron transport, and<br />

ATP formation, and also produces an ADP ribosylation of the GAPDH. This process<br />

results in acute endothelial dysfunction in diabetic blood vessels that contributes to the<br />

development of diabetic complications. NF-kB activation also induces a pro-inflammatory<br />

condition and adhesion molecule over-expression. All these alterations produce the final<br />

picture of diabetic complications.<br />

transport, and ATP formation and also produces an ADP ribosylation of the<br />

GAPDH. 59 These processes result in acute endothelial dysfunction in diabetic<br />

blood vessels that, convincingly, contributes to the development of CVD. 59 These<br />

pathways are summarized in Figure 8.1.<br />

Several indirect and direct approaches support the concept that acute hyperglycemia<br />

works through the production of oxidative and nitrosative stress. Indirect<br />

evidence is obtained through the use of antioxidants. The fact that antioxidants


130 Oxidative Stress and Inflammatory Mechanisms<br />

can hinder some of the effects acutely induced by hyperglycemia–endothelial<br />

dysfunction, 36,60,61 activation of coagulation, 44 plasmatic increase of ICAM-1, 53<br />

and interleukins 57 suggests that the action of acute hyperglycemia is mediated by<br />

the production of free radicals.<br />

Direct evidence is linked to estimates of the effects of acute hyperglycemia<br />

on oxidative stress markers. It has been reported that during oral glucose challenge,<br />

a reduction of the antioxidant defenses was observed. 62–64 This effect can<br />

be observed in other physiologic situations related to eating a meal. 65 The role<br />

of hyperglycemia is highlighted by the results of a study involving two different<br />

meals resulting in two different levels of post-prandial hyperglycemia. The greater<br />

drop in antioxidant activity was linked with the higher levels of hyperglycemia. 34<br />

The evidence that LDLs are more prone to oxidation in the post-prandial phase<br />

in diabetics matches these results. 33 Even in this situation, higher levels of hyperglycemia<br />

are matched with a greater oxidation of LDLs. 34 Finally, the evidence<br />

that managing post-prandial hyperglycemia can reduce post-prandial generation<br />

of endothelial dysfunction 66 and oxidative and nitrosative stress 67 strongly supports<br />

this hypothesis.<br />

Interesting and new data are available on the possible generation of nitrosative<br />

stress during post-prandial hyperglycemia. The simultaneous over-generation of<br />

NO and superoxide favors the production of a toxic reaction product, the peroxynitrite<br />

anion. 68 The peroxynitrite anion is cytotoxic because it oxidizes sulfhydryl<br />

groups in proteins, initiates lipid peroxidation, and nitrates amino acids such as<br />

tyrosine which affects many signal transduction pathways. 68 The production of<br />

peroxynitrite can be indirectly inferred by the presence of nitrotyrosine, 68 and it<br />

has recently been reported that nitrotyrosine is an independent predictor of CVD. 69<br />

Several pieces of evidence support a direct role of hyperglycemia in favoring<br />

nitrotyrosine over-generation. Nitrotyrosine formation was detected in the artery<br />

walls of monkeys during hyperglycemia 70 and also in the plasma of healthy<br />

subjects during hyperglycemic clamp 71 or OGTT. 72,73 Hyperglycemia was also<br />

accompanied by nitrotyrosine deposition in perfused working hearts from rats,<br />

and was reasonably related to unbalanced production of NO and superoxide<br />

through iNOS over-expression. 74 Nitrotyrosine formation is followed by the<br />

development of endothelial dysfunction in both healthy subjects 71,72 and in coronaries<br />

of perfused hearts. 74 This effect is not surprising because it has been<br />

shown that nitrotyrosine can also be directly harmful to endothelial cells. 75<br />

Dyslipidemia also is a recognized risk factor for cardiovascular disease in<br />

diabetes 76 and post-prandial hyperlipidemia contributes to this risk. 77 In non-obese<br />

type 2 diabetes patients with moderate fasting hypertriglyceridemia, atherogenic<br />

lipoprotein profiles were amplified in the post-prandial state. 78 Such observations<br />

have raised the question of whether post-prandial hyperlipidemia, which rises<br />

concomitantly with post-prandial hyperglycemia, is the true risk factor. 79 Evidence<br />

suggests that post-prandial hypertriglyceridemia and hyperglycemia independently<br />

induce endothelial dysfunction through oxidative stress. 80 It is now<br />

well recognized that endothelial dysfunction is one of the first stages and<br />

earliest markers in the development of cardiovascular disease. 81 Recent studies


Dysfunction, Stress, and Inflammation in Type 2 Diabetes 131<br />

demonstrate both independent and cumulative effects of post-prandial hypertriglyceridemia<br />

and hyperglycemia on endothelial function, with oxidative stress as<br />

the common mediator. 72,73 This lends credence to the idea of a direct atherogenic<br />

role for post-prandial hyperglycemia independent of the role of lipids.<br />

CONCLUSIONS<br />

The evidence described proves that hyperglycemia can acutely induce alterations<br />

of normal human homeostasis. It should be noted that acute increases of glucose<br />

serum level cause alterations in both healthy (normoglycemic) subjects and also<br />

in diabetic subjects who have basic hyperglycemia. On the basis of this evidence<br />

we can hypothesize that the acute effects of glucose serum levels can add to those<br />

produced by chronic hyperglycemia, thus contributing to a final picture of complicated<br />

diabetes. The precise relevance of this phenomenon is not exactly comprehensible<br />

and quantifiable at the moment, but based on the tendency to rapid<br />

variations of hyperglycemia in the lives of diabetic patients, primarily in the<br />

post-prandial phase, it is proper to think that it may exert an influence on the<br />

onset of complications. Epidemiological studies 3 and preliminary intervention<br />

studies seem to support this hypothesis. 27–30<br />

DCCT in type 1 diabetes 82 and UKPDS in type 2 diabetes 5 both attest to the<br />

importance of long-term glycemic control through HbA1c for the prevention of<br />

complications. However, the authors of the DCCT pointed out further that HbA1c<br />

only is not a sufficient parameter to explain the onset of such complications and<br />

suggested that post-prandial hyperglycemic excursions could reasonably favor<br />

the onset of diabetic complications. 82<br />

Evidence shows that post-prandial glucose serum level is the major determinant<br />

of HbA1c level after mean daily blood glucose 83–86 and that reducing<br />

post-prandial hyperglycemia significantly reduces HbA1c levels in type 2 diabetic<br />

patients. 87,88 On the basis of this evidence, it seems obvious that if post-prandial<br />

hyperglycemia is important to determine the level of HbA1c, which is fundamental<br />

to determining the extent of the risk of diabetic complications, it can be<br />

supposed that post-prandial glucose serum level will favor it as much.<br />

Evidence also suggests that post-prandial excursions of blood glucose may<br />

be involved in the development of diabetic complications, particularly cardiovascular<br />

complications, but are not the only factors. 89,90 Many questions remain<br />

unanswered regarding the definition of post-prandial glucose and, perhaps most<br />

importantly, whether post-prandial hyperglycemia has a unique role in the pathogenesis<br />

of diabetic vascular complications and should be a specific target of<br />

therapy.<br />

However, this alarmingly suggestive body of evidence for a harmful effect<br />

of post-prandial hyperglycemia on diabetic complications has been sufficient to<br />

influence guidelines from key professional bodies including the World Health<br />

Organization, 91 the American Diabetes Association, 92 the American College of<br />

Endocrinology, 93 the International Diabetes Federation, 94 the Canadian Diabetes


132 Oxidative Stress and Inflammatory Mechanisms<br />

Association, 95 and more recently a large task force of European scientific societies<br />

focusing on cardiovascular disease. 96<br />

Therefore the real question, as recently underlined also by the American<br />

Diabetes Association, 97 seems to be that “because CVD is the major cause of<br />

morbidity and mortality in patients with diabetes, and in type 2 diabetes in<br />

particular, understanding the impact on CVD events of treatment directed at<br />

specifically lowering post-prandial glucose is crucial.” Future studies must be<br />

designed specifically to evaluate this fundamental issue that may significantly<br />

change the therapeutic approach to diabetes.<br />

REFERENCES<br />

1. Kannel WB, McGee DL. Diabetes and cardiovascular diseases: the Framingham<br />

Study. JAMA 241, 2035, 1979.<br />

2. Laakso M. Hyperglycemia and cardiovascular disease in type 2 diabetes. Diabetes<br />

48, 937, 1999.<br />

3. Bonora E and Muggeo M. Post-prandial blood glucose as a risk factor for cardiovascular<br />

disease in type II diabetes: epidemiological evidence. Diabetologia 44,<br />

2107, 2001.<br />

4. Stratton IM et al. Association of glycaemia with macrovascular and microvascular<br />

complications of type 2 diabetes (UKPDS 35): prospective observational study.<br />

Br Med J 321, 405, 2000.<br />

5. United Kingdon Prospective Diabetes Study Group. Intensive blood-glucose control<br />

with sulphonylureas or insulin compared with conventional treatment and risk<br />

of complications in patients with type 2 diabetes (UKPDS 33). Lancet 352, 837,<br />

1998.<br />

6. United Kingdom Prospective Diabetes Study Group. Effect of intensive bloodglucose<br />

control with metformin on complications in overweight patients with type<br />

2 diabetes (UKPDS 34). Lancet 352, 854, 1998.<br />

7. Vinik AI and Jenkins DJ. Dietary fiber in management of diabetes. Diabetes Care<br />

11, 160, 1988.<br />

8. Wolever TMS et al. Variation of post-prandial plasma glucose, palatability, and<br />

symptoms associated with a standardized mixed test meal versus 75 g oral glucose.<br />

Diabetes Care 21, 336, 1998.<br />

9. de Vegt F et al. Hyperglycaemia is associated with all-cause and cardiovascular<br />

mortality in the Hoorn population: the Hoorn study. Diabetologia 42, 926, 1999.<br />

10. Donahue RP, Abbott RD, Reed DM, and Yano K. Post-challenge glucose concentration<br />

and coronary heart disease in men of Japanese ancestry: Honolulu Heart<br />

Program. Diabetes 36, 689, 1987.<br />

11. Lowe LP, Liu K, Greenland P, Metzger BE, Dyer AR, and Stamler J. Diabetes,<br />

asymptomatic hyperglycemia, and 22-year mortality in black and white men: the<br />

Chicago Heart Association Detection Project in Industry Study. Diabetes Care<br />

20, 163, 1997.<br />

12. The DECODE study group on behalf of the European Diabetes Epidemiology<br />

Group. Glucose tolerance and mortality: comparison of WHO and American<br />

Diabetes Association diagnostic criteria. Lancet 354, 617, 1999.


Dysfunction, Stress, and Inflammation in Type 2 Diabetes 133<br />

13. Coutinho M, Gerstein HC, Wang Y, and Yusuf S. The relationship between glucose<br />

and incident cardiovascular events. a metaregression analysis of published data<br />

from 20 studies of 95,783 individuals followed for 12.4 years. Diabetes Care 22,<br />

233, 1999.<br />

14. Balkau B et al. High blood glucose concentration is a risk factor for mortality in<br />

middle-aged nondiabetic men: 20-year follow-up in the Whitehall Study, the Paris<br />

Prospective Study, and the Helsinki Policemen Study. Diabetes Care 21, 360,<br />

1998.<br />

15. Hanefeld M et al. The DIS Group. Risk factors for myocardial infarction and<br />

death in newly detected NIDDM: Diabetes Intervention Study, 11-year follow-up.<br />

Diabetologia 39, 1577, 1996.<br />

16. Hanefeld M et al. Post-prandial plasma glucose is an independent risk factor for<br />

increased carotid intima-media thickness in non-diabetic individuals. Atherosclerosis<br />

144, 229, 1999.<br />

17. Temelkova-Kurktschiev TS et al. Post-challenge plasma glucose and glycemic<br />

spikes are more strongly associated with atherosclerosis than fasting glucose and<br />

glycosylated hemoglobin level. Diabetes Care 23, 1830, 2000.<br />

18. Bellodi G et al. Hyperglycemia and prognosis of acute myocardial infarction in<br />

patients without diabetes mellitus. Am J Cardiol 64, 885, 1989.<br />

19. O’Sullivan JJ, Conroy RM, Robinson K, Hickey N, and Mulcahy R. In-hospital<br />

prognosis of patients with fasting hyperglycemia after first myocardial infarction.<br />

Diabetes Care 14, 758, 1991.<br />

20. Gray CS et al. The prognostic value of stress hyperglycaemia and previously<br />

unrecognized diabetes in acute stroke. Diabet Med 4, 237, 1987.<br />

21. Gray CS, French JM, Bates D, Cartlidge NE, Venables GS, and James OF.<br />

Increasing age, diabetes mellitus and recovery from stroke. Postgrad Med J 65,<br />

720, 1989.<br />

22. Capes SE, Hunt D, Malmberg K, and Gerstein HC. Stress hyperglycaemia and<br />

increased risk of death after myocardial infarction in patients with and without<br />

diabetes: a systematic overview. Lancet 355, 773, 2000.<br />

23. Malmberg K, Norhammar A, Wedel H, and Ryden L. Glycometabolic state at<br />

admission. important risk marker of mortality in conventionally treated patients<br />

with diabetes mellitus and acute myocardial infarction: long-term results from the<br />

diabetes and insulin-glucose infusion in acute myocardial infarction (DIGAMI)<br />

study. Circulation 99, 2626, 1999.<br />

24. Marfella R, Nappo F, De Angelis L, Siniscalchi M, Rossi F, and Giugliano D.<br />

The effect of acute hyperglycaemia on QTc duration in healthy man. Diabetologia<br />

43, 571, 2000.<br />

25. Gokhroo R and Mittal SR. Electrocardiographic correlates of hyperglycemia in<br />

acute myocardial infarction. Int J Cardiol 22, 267, 1989.<br />

26. Chiasson JL, Josse RG, Gomis R, Hanefeld M, Karasik A, and Laakso M. Acarbose<br />

for prevention of type 2 diabetes mellitus: STOP-NIDDM randomised trial.<br />

Lancet 359, 2072, 2002.<br />

27. Chiasson JL, Josse RG, Gomis R, Hanefeld M, Karasik A, and Laakso M. Acarbose<br />

treatment and the risk of cardiovascular disease and hypertension in patients<br />

with impaired glucose tolerance: STOP-NIDDM trial. JAMA 290, 486, 2003.<br />

28. Hanefeld M et al. Acarbose slows progression of intima-media thickness of the<br />

carotid arteries in subjects with impaired glucose tolerance. Stroke 35, 1073, 2004.


134 Oxidative Stress and Inflammatory Mechanisms<br />

29. Hanefeld M, Cagatay M, Petrowitsch T, Neuser D, Petzinna D, and Rupp M.<br />

Acarbose reduces the risk for myocardial infarction in type 2 diabetic patients:<br />

meta-analysis of seven long-term studies. Eur Heart J 25, 10, 2004.<br />

30. Esposito K, Giugliano D, Nappo F, and Marfella R. Regression of carotid atherosclerosis<br />

by control of post-prandial hyperglycemia in type 2 diabetes mellitus.<br />

Circulation 29, 2978, 2004.<br />

31. Tsai EC, Hirsch IB, Brunzell JD, and Chait A. Reduced plasma peroxyl radical<br />

trapping capacity and increased susceptibility of LDL to oxidation in poorly<br />

controlled IDDM. Diabetes 43, 1010, 1994.<br />

32. Jenkins AJ et al. LDL from patients with well-controlled IDDM is not more<br />

susceptible to in vitro oxidation. Diabetes 45, 762, 1996.<br />

33. Diwadkar VA, Anderson JW, Bridges SR, Gowri MS, and Oelgten PR. Post-prandial<br />

low density lipoproteins in type 2 diabetes are oxidized more extensively than<br />

fasting diabetes and control samples. Proc Soc Exp Biol Med 222, 178, 1999.<br />

34. Ceriello A et al. Meal-induced oxidative stress and low-density lipoprotein oxidation<br />

in diabetes: possible role of hyperglycemia. Metabolism 48, 1503, 1999.<br />

35. Jorgensen RG, Russo L, Mattioli L, and Moore WV. Early detection of vascular<br />

dysfunction in type I diabetes. Diabetes 37, 292, 1988.<br />

36. Marfella R et al. Glutathione reverses systemic hemodynamic changes by acute<br />

hyperglycemia in healthy subjects. Am J Physiol 268, E1167, 1995.<br />

37. Kawano H et al. Hyperglycemia rapidly suppresses flow-mediated endotheliumdependent<br />

vasodilation of brachial artery. J Am Coll Cardiol 34, 146, 1999.<br />

38. Giugliano D et al. Vascular effects of acute hyperglycemia in humans are reversed<br />

by L-arginine: evidence for reduced availability of nitric oxide during hyperglycemia.<br />

Circulation 95, 1783, 1997.<br />

39. Shige H et al. Endothelium-dependent flow-mediated vasodilation in the post-prandial<br />

state in type 2 diabetes mellitus. Am J Cardiol 84, 1272, 1999.<br />

40. Ceriello A. Coagulation activation in diabetes mellitus: the role of hyperglycaemia<br />

and therapeutic prospects. Diabetologia 36, 1119, 1993.<br />

41. Jones RL and Peterson CM. Reduced fibrinogen survival in diabetes mellitus: a<br />

reversible phenomenon. J Clin Invest 63, 485, 1979.<br />

42. Jones RL. Fibrinopeptide A in diabetes mellitus: relation to levels of blood glucose,<br />

fibrinogen disappearance, and hemodynamic changes. Diabetes 34, 836, 1985.<br />

43. Ceriello A et al. Hyperglycemia may determine fibrinopeptide A plasma level<br />

increase in humans. Metabolism 38, 1162, 1989.<br />

44. Ceriello A et al. Hyperglycemia-induced thrombin formation in diabetes: possible<br />

role of oxidative stress. Diabetes 44, 924, 1995.<br />

45. Ceriello A, Giugliano D, Quatraro A, Dello Russo P, and Torella R. Blood glucose<br />

may condition factor VII levels in diabetic and normal subjects. Diabetologia 31,<br />

889, 1988.<br />

46. Sakamoto T et al. Rapid change of platelet aggregability in acute hyperglycemia:<br />

detection by a novel laser-light scattering method. Thromb Haemost 83, 475, 2000.<br />

47. Ceriello A et al. Post-meal coagulation activation in diabetes mellitus: effect of<br />

acarbose. Diabetologia 39, 469, 1996.<br />

48. Ruosladti E. Integrins. J Clin Invest 187, 1, 1991.<br />

49. Lopes-Virella MF and Virella G. Immune mechanism of atherosclerosis in diabetes<br />

mellitus. Diabetes 41, Suppl. 2, 86, 1992.<br />

50. Blann AD and McCollum CN. Circulating endothelial cell/leukocyte adhesion<br />

molecules in atherosclerosis. Thromb Haemost 72, 151, 1994.


Dysfunction, Stress, and Inflammation in Type 2 Diabetes 135<br />

51. Cominacini L et al. Elevated levels of soluble E-selectin in patients with IDDM<br />

and NIDDM: relation to metabolic control. Diabetologia 38, 1122, 1995.<br />

52. Ceriello A et al. Increased circulating ICAM-1 levels in type 2 diabetic patients:<br />

possible role of metabolic control and oxidative stress. Metabolism 45, 498, 1996.<br />

53. Ceriello A et al. Hyperglycemia-induced circulating ICAM-1 increase in diabetes<br />

mellitus: possible role of oxidative stress. Horm Metab Res 30, 146, 1998.<br />

54. Marfella R et al. Circulating adhesion molecules in humans: role of hyperglycemia<br />

and hyperinsulinemia. Circulation 101, 2247, 2000.<br />

55. Plutzky J. Inflammation in atherosclerosis and diabetes mellitus. Rev Endocr<br />

Metab Disord 5, 255, 2004.<br />

56. Esposito K et al. Inflammatory cytokine concentrations are acutely increased by<br />

hyperglycemia in humans: role of oxidative stress. Circulation 106, 2067, 2002.<br />

57. Nappo F et al. Post-prandial endothelial activation in healthy subjects and in type<br />

2 diabetic patients: role of fat and carbohydrate meals. J Am Coll Cardiol 39,1145,<br />

2002.<br />

58. Brownlee M. Biochemistry and molecular cell biology of diabetic complications.<br />

Nature 414, 813, 2001.<br />

59. Ceriello A. New insights on oxidative stress and diabetic complications may lead<br />

to a “causal” antioxidant therapy. Diabetes Care 26, 1589, 2003.<br />

60. Title LM, Cummings PM, Giddens K, and Nassar BA. Oral glucose loading<br />

acutely attenuates endothelium-dependent vasodilation in healthy adults without<br />

diabetes: effect prevented by vitamins C and E. J Am Coll Cardiol 36, 2185, 2000.<br />

61. Beckman JA, Goldfine AB, Gordon MB, and Creager MA. Ascorbate restores<br />

endothelium-dependent vasodilation impaired by acute hyperglycemia in humans.<br />

Circulation 103, 1618, 2001.<br />

62. Ceriello A et al. Antioxidant defenses are reduced during oral glucose tolerance<br />

test in normal and non-insulin dependent diabetic subjects. Eur J Clin Invest 28,<br />

329, 1998.<br />

63. Tessier D, Khalil A, and Fulop T. Effects of an oral glucose challenge on free<br />

radical/antioxidant balance in an older population with type II diabetes. J Gerontol<br />

54, 541, 1999.<br />

64. Konukoglu D, Hatemi H, Ozer EM, Gonen S, and Akcay T. The erythrocyte<br />

glutathione levels during oral glucose tolerance test. J Endocrinol Invest 20, 471,<br />

1997.<br />

65. Ceriello A et al. Meal-generated oxidative stress in type 2 diabetic patients.<br />

Diabetes Care 21, 1529, 1998.<br />

66. Ceriello A et al. The post-prandial state in type 2 diabetes and endothelial dysfunction:<br />

effects of insulin aspart. Diabet Med 21, 171, 2004.<br />

67. Ceriello A et al. Role of hyperglycemia in nitrotyrosine post-prandial generation.<br />

Diabetes Care 25, 1439, 2002.<br />

68. Beckman JS and Koppenol WH. Nitric oxide, superoxide, and peroxynitrite: the<br />

good, the bad, and ugly. Am J Physiol 271, C1424, 1996.<br />

69. Shishehbor MH et al. Association of nitrotyrosine levels with cardiovascular<br />

disease and modulation by statin therapy. JAMA 289, 1675, 2003.<br />

70. Pennathur S, Wagner JD, Leeuwenburgh C, Litwak KN, and Heinecke JW. A<br />

hydroxyl radical-like species oxidizes cynomolgus monkey artery wall proteins<br />

in early diabetic vascular disease. J Clin Invest 107, 853, 2001.


136 Oxidative Stress and Inflammatory Mechanisms<br />

71. Marfella R, Quagliaro L, Nappo F, Ceriello A, and Giugliano D. Acute hyperglycemia<br />

induces an oxidative stress in healthy subjects (letter). J Clin Invest 108,<br />

635, 2001.<br />

72. Ceriello A et al. Evidence for an independent and cumulative effect of post-prandial<br />

hypertriglyceridemia and hyperglycemia on endothelial dysfunction and oxidative<br />

stress generation: effects of short- and long-term simvastatin treatment.<br />

Circulation 106, 1211, 2002.<br />

73. Ceriello A et al. Effect of post-prandial hypertriglyceridemia and hyperglycemia<br />

on circulating adhesion molecules and oxidative stress generation and the possible<br />

role of simvastatin treatment. Diabetes 53, 701, 2004.<br />

74. Ceriello A et al. Acute hyperglycemia induces nitrotyrosine formation and apoptosis<br />

in perfused heart from rat. Diabetes 51, 1076, 2002.<br />

75. Mihm MJ, Jing L, and Bauer JA. Nitrotyrosine causes selective vascular endothelial<br />

dysfunction and DNA damage. J Cardiovasc Pharmacol 36, 182, 2000.<br />

76. Taskinen MR, Lahdenpera S, and Syvanne M. New insights into lipid metabolism<br />

in non-insulin-dependent diabetes mellitus. Ann Med 28, 335, 1996.<br />

77. Karpe F, de Faire U, Mercuri M, Bond MG, Hellenius ML, and Hamsten A.<br />

Magnitude of alimentary lipemia is related to intima-media thickness of the<br />

common carotid artery in middle-aged men. Atherosclerosis 141, 307, 1998.<br />

78. Cavallero E, Dachet C, Neufcou D, Wirquin E, Mathe D, and Jacotot B. Post-prandial<br />

amplification of lipoprotein abnormalities in controlled type II diabetic subjects:<br />

relationship to post-prandial lipemia and C-peptide/glucagon levels.<br />

Metabolism 43, 270, 1994.<br />

79. Heine RJ and Dekker JM. Beyond post-prandial hyperglycemia: metabolic factors<br />

associated with cardiovascular disease. Diabetologia 45, 461, 2002.<br />

80. Ceriello A and Motz E. Is oxidative stress the pathogenic mechanism underlying<br />

insulin resistance, diabetes, and cardiovascular disease? The common soil hypothesis<br />

revisited. Arterioscler Thromb Vasc Biol 24, 816, 2004.<br />

81. De Caterina R. Endothelial dysfunctions. common denominators in vascular disease.<br />

Curr Opin Lipidol 11, 923, 2000.<br />

82. The Diabetes Control and Complications Trial Research Group. The relationship<br />

of glycemic exposure (HbA1c) to the risk of development and progression of<br />

retinopathy in the Diabetes Control and Complications Trial. Diabetes 44, 968,<br />

1995.<br />

83. Avignon A, Radauceanu A, and Monnier L. Nonfasting plasma glucose is a better<br />

marker of diabetic control than fasting plasma glucose in type 2 diabetes. Diabetes<br />

Care 20, 1822, 1997.<br />

84. Soonthornpun S, Rattarasarn C, Leelawattana R, and Setasuban W. Post-prandial<br />

plasma glucose: a good index of glycemic control in type 2 diabetic patients<br />

having near-normal fasting glucose levels. Diabetes Res Clin Pract 46, 23, 1999.<br />

85. Monnier L, Lapinski H, and Colette C. Contributions of fasting and post-prandial<br />

plasma glucose increments to the overall diurnal hyperglycemia of type 2 diabetic<br />

patients: variations with increasing levels of HbA(1c). Diabetes Care 26, 881,<br />

2003.<br />

86. Rohlfing CL et al. Defining the relationship between plasma glucose and HbA(1c):<br />

analysis of glucose profiles and HbA(1c) in the Diabetes Control and Complications<br />

Trial. Diabetes Care 25, 275, 2002.<br />

87. Bastyr EJ 3rd et al. Therapy focused on lowering post-prandial glucose, not fasting<br />

glucose, may be superior for lowering HbA1c. Diabetes Care 23, 1236, 2000.


Dysfunction, Stress, and Inflammation in Type 2 Diabetes 137<br />

88. Home PD, Lindholm A, and Hylleberg B, Round P. Improved glycemic control<br />

with insulin aspart. a multicenter randomized double-blind crossover trial in type<br />

1 diabetic patients. Diabetes Care 21, 1904, 1998.<br />

89. Shichiri M, Kishikawa H, Ohkubo Y, and Wake N. Long-term results of Kumamoto<br />

Study on optimal diabetes control in type 2 diabetic patients. Diabetes Care 23,<br />

Suppl. 2, B21, 2000.<br />

90. Singleton JR, Smith AG, Russell JW, and Feldman EL. Microvascular complications<br />

of impaired glucose tolerance. Diabetes 52, 2867, 2003.<br />

91. World Health Organization. Definition, diagnosis and classification of diabetes<br />

mellitus and its complications. Part 1. Geneva, 1999.<br />

92. American Diabetes Association. Standards of medical care in diabetes. Diabetes<br />

Care 27, Suppl.1, S15, 2004.<br />

93. American College of Endocrinology. Consensus statement on guidelines for glycemic<br />

control. Endocr Pract 8, 5, 2002.<br />

94. Alberti KG and Gries FA. Management of non-insulin-dependent diabetes mellitus<br />

in Europe: a consensus view. Diabet Med 5, 275, 1988.<br />

95. Canadian Diabetes Association, Clinical Practice Guidelines Expert Committee.<br />

Clinical practice guidelines for the prevention and management of diabetes in<br />

Canada. Can J Diabetes Suppl. 2, 27, 2003.<br />

96. De Backer G et al. Third Joint Task Force of European and Other Societies on<br />

Cardiovascular Disease Prevention in Clinical Practice. European guidelines on<br />

cardiovascular disease prevention in clinical practice. Eur Heart J 24, 1601, 2003.<br />

97. American Diabetes Association. Postprandial blood glucose. Diabetes Care 24,<br />

775, 2001.


9<br />

CONTENTS<br />

Obesity and<br />

Inflammation:<br />

Implications for<br />

Atherosclerosis<br />

John Alan Farmer<br />

Introduction .......................................................................................................139<br />

Obesity and Cardiovascular Risk......................................................................141<br />

Clinical Markers of Inflammation ....................................................................143<br />

C-Reactive Protein...................................................................................144<br />

Leptin .....................................................................................................146<br />

Adiponectin..............................................................................................147<br />

Tumor Necrosis Factor-α ........................................................................148<br />

Hemostatic Parameters ............................................................................150<br />

Resistin ....................................................................................................151<br />

Other Inflammatory Markers...................................................................152<br />

Atherosclerosis and Inflammation ....................................................................152<br />

Obesity and Inflammation.................................................................................154<br />

Modification of Inflammation in Obesity.........................................................156<br />

C-Reactive Protein...................................................................................157<br />

Tumor Necrosis Factor-α ........................................................................159<br />

Hemostatic Factors ..................................................................................159<br />

Summary ...........................................................................................................160<br />

References .........................................................................................................160<br />

INTRODUCTION<br />

Age-adjusted mortality rates for cardiovascular disease have been significantly<br />

declining in the United States over the past 20 years. The encouraging reduction<br />

is due to a complex interaction of continued improvements in diagnostic imaging,<br />

medical therapy, and surgical techniques. The pathogenesis of atherosclerosis is<br />

139


140 Oxidative Stress and Inflammatory Mechanisms<br />

complex and best regarded as a syndrome currently lacking a unifying hypothesis<br />

that explains all aspects of the initiation and progression of vascular disease.<br />

However, the recognition and modification of potential cardiovascular risk factors<br />

that have been statistically related to coronary disease (in particular factors whose<br />

modification has been demonstrated to reduce clinical event rates in prospective<br />

controlled clinical trials) have gained credence as a means to reduce the risk from<br />

atherosclerosis.<br />

The improvements in cardiovascular morbidity and mortality rates are threatened<br />

by changing patterns in obesity and diabetes. The incidence and prevalence<br />

of obesity are significantly increasing in the United States and represent a major<br />

health hazard due to the independent risks directly related to increased body mass<br />

index per se in addition to the associated potential risks for the development of<br />

diabetes, dyslipidemia, and hypertension. 1<br />

Individuals may be classified as normal, overweight, obese, or morbidly obese<br />

on the basis of body mass index (weight in kilograms over body surface area in<br />

square meters). Normality determined by body mass index is considered below<br />

25. A subject may be classified as overweight if the body mass index falls between<br />

25 and 30. Obesity is defined as a body mass index in excess of 30 and morbid<br />

obesity requires a body mass index above 40.<br />

Recent epidemiologic data concerning the prevalence of obesity in the United<br />

States utilized the National Health and Nutrition Education Survey (NHANES). 2<br />

The prevalence of subjects who qualified as overweight or obese in the period<br />

between years 2003 and 2004 was compared to an earlier phase of the survey<br />

that analyzed subjects in 1999 and 2000. The prevalence of obesity increased<br />

from 27.5 to 31.1% in male subjects between the designated time periods. In<br />

comparison, the prevalence of obesity in women remained relatively constant at<br />

approximately 33%. The prevalence of overweight and obese children and adolescents<br />

was disturbingly high (17.1%). The overall prevalence of obesity in the<br />

general United States population was determined to be 32.2%.<br />

The increasing prevalence of obesity and related conditions has led to significant<br />

health and economic concerns due to the dramatic impacts on both direct<br />

and indirect health care costs. The total economic costs of obesity in the United<br />

States are estimated to account for expenditures of approximately $60 billion. 3<br />

Although the interpretation of the actuarial data became controversial after intense<br />

scrutiny, the Centers for Disease Control previously estimated that approximately<br />

400,000 deaths per year can be directly attributed to obesity. Furthermore, the<br />

health risks associated with increased body weight may overtake the quantitative<br />

impact of tobacco as a contributor to overall cardiac risk assuming the present<br />

trends continue.<br />

Obesity had been previously regarded simply as a passive deposition of excess<br />

energy stored in adipose tissue. However, an increasing body of clinical and<br />

experimental evidence has implicated adipose tissue as a highly active endocrine<br />

organ that is intimately involved in a variety of clinically important metabolic<br />

pathways that subsequently increase the risk for atherosclerosis. Despite a clear<br />

univariate statistical correlation between increased body mass index and total or


Obesity and Inflammation: Implications for Atherosclerosis 141<br />

cardiovascular mortality, the independent cause-and-effect relationship is difficult<br />

to establish with definite certainty due to multiple potential confounding factors<br />

including degree of physical activity, regional distribution of fat (truncal versus<br />

gluteal), socioeconomic status, and associated metabolic disorders such as hypertension,<br />

hyperlipidemia, and diabetes. Age may also play a role in the impact of<br />

obesity on mortality; the statistical relationship may attenuate over time and<br />

become less prominent in older age groups. 4 This review will focus on the role<br />

of the adipocyte in inflammation and the potential impact on cardiovascular risk.<br />

OBESITY <strong>AND</strong> CARDIOVASCULAR RISK<br />

Obesity has been statistically associated with profound reductions in lifespan in<br />

both men and women. For example, a 20-year-old Caucasian man with a body<br />

mass index in excess of 45 is estimated to lose 13 years of life due to medical<br />

problems directly related to obesity. 5 Cardiovascular disease (including cerebrovascular<br />

accident and sudden cardiac death) is a major component of the<br />

increased mortality associated with obesity.<br />

However, significant clinical and statistical problems are encountered in the<br />

quantification of the impact of obesity on cardiovascular mortality due to the<br />

common association of a variety of intimately interrelated and frequently coexistent<br />

metabolic conditions such as diabetes, dyslipidemia, and hypertension.<br />

Additionally, individuals who are classified as overweight purely on the basis<br />

of body mass index may be relatively healthy if free of associated metabolic<br />

disorders.<br />

The concept that overweight or obese individuals who are physically fit and<br />

do not have associated metabolic conditions may not demonstrate increased<br />

cardiovascular risk has been proposed although considerable controversy exists<br />

due to methodological problems and the lack of large-scale prospective controlled<br />

clinical trials. 6 However, it is clear that a lean physically fit individual has the<br />

lowest overall risk of developing cardiovascular disease. The presence and quantitative<br />

contributions of lifestyle, diet, exercise patterns, and degrees of physical<br />

fitness are difficult to determine in epidemiologic studies and may potentially<br />

confound the interpretation of observational data.<br />

The duration and intensity of physical activity are also significant determinants<br />

of body weight, glucose tolerance, and blood pressure but the quantitation<br />

of physical fitness is frequently problematic in large scale observational studies.<br />

The relationship of cardiovascular fitness, mortality, and obesity is conflicting.<br />

Studies have demonstrated that the highest relative risk of all-cause mortality was<br />

evident in obese individuals who were physically unfit. 7 However, subjects who<br />

were of normal weight and physically unfit appeared to have higher relative risks<br />

of mortality when compared to obese individuals with optimal cardio-respiratory<br />

fitness. The concept of the metabolically healthy obese (MHO) individual has<br />

been proposed as a means to eliminate confounding findings but is difficult to<br />

define and has not traditionally been evaluated in epidemiologic studies.


142 Oxidative Stress and Inflammatory Mechanisms<br />

Multiple observational trials performed in men have linked physical fitness<br />

and health benefits with resultant reductions in risk from cardiovascular disease. 8<br />

The benefits from cardiovascular fitness have also been demonstrated in women.<br />

However, conflicting results relative to the contributions of physical fitness and<br />

obesity may be gender-related. The relationship of physical fitness, obesity, and<br />

cardiovascular events in roughly 1000 women was determined over a 4-year<br />

observational period. 9 Women who were demonstrated to have low levels of<br />

physical activity independent of obesity were significantly more likely to have<br />

major adverse cardiovascular events during the 48-month follow-up period. The<br />

results were interpreted to imply that physical fitness and metabolic health may<br />

be more important than obesity per se for the determination of cardiovascular<br />

risk in women. The presence of conflicting data has generated a significant debate<br />

concerning the relative role of “fitness versus fatness” in cardiovascular and total<br />

mortality and has not been completely resolved.<br />

Diabetes mellitus has been termed a coronary equivalent due to the high<br />

risks of developing cardiac events in diabetic subjects. The role of physical fitness<br />

and alteration of body mass index as a means to reduce the risk of developing<br />

diabetes has been recently analyzed. 10 The relative contribution of body mass<br />

index as a means to predict the subsequent risk of development of type II diabetes<br />

appeared to be stronger than the degree of physical activity. Additionally,<br />

improved physical activity appeared to have little effect upon the relationship<br />

between changes in body mass index and development of diabetes. However,<br />

definite clinical endpoints were not evaluated and the method employed to report<br />

physical activity may not have been optimal. Thus the relative roles of physical<br />

activity, metabolic health, and obesity are difficult to separate in observational<br />

studies and represent significant problems for related risk factors such as dyslipidemia<br />

and hypertension.<br />

Despite the methodological difficulties in the analysis of the quantitative<br />

independent contribution of obesity to global cardiovascular risk, it is clear that<br />

metabolic factors associated with obesity play a major role in the subsequent<br />

incidence of cardiac events. The term metabolic syndrome describes a constellation<br />

of cardiovascular risk factors (including truncal obesity) that appear to share<br />

common metabolic pathways as a means to identify and target high risk individuals<br />

for aggressive multifactorial therapeutic interventions. The National Cholesterol<br />

Education Program has established diagnostic criteria and requires any three<br />

of the five following risk factors 11 to be present in order to make the diagnosis<br />

of metabolic syndrome (Figure 9.1):<br />

1. Abdominal obesity is defined as waist circumference measured at the<br />

umbilicus and is considered to be positive if in excess of 40" in men<br />

and 35" in women.<br />

2. Fasting triglycerides must be in excess of 150 mg per deciliter.<br />

3. Fasting glucose: blood sugar must be in excess of 110 mg per deciliter.


Obesity and Inflammation: Implications for Atherosclerosis 143<br />

Fasting<br />

Glucose<br />

>110 mg<br />

HDL<br />

Men 35<br />

inches<br />

Blood<br />

Pressure<br />

>130<br />

>85<br />

FIGURE 9.1 Clinical diagnosis of the metabolic syndrome.<br />

Triglyceride<br />

>150 mg/dl<br />

4. HDL cholesterol must be less than 40 mg per deciliter in men and less<br />

than 50 mg per deciliter in women.<br />

5. Blood pressure must be in excess of 130 mm of mercury systolic and<br />

85 mm of mercury diastolic.<br />

The clinical utility of the concept of the metabolic syndrome as a specific disease<br />

entity has been challenged by the American Diabetes Association and the European<br />

Association for the Study of Diabetes. 12 However, the use of the syndrome<br />

term does not necessarily imply a specific disease entity but refers to a grouping<br />

of clinical symptoms or conditions that appear to coexist in individuals out of<br />

proportion to the prevalence in the general community. The major cardiovascular<br />

risk factors share multiple metabolic pathways that have been linked to the process<br />

of atherosclerosis. The Common Soil Hypothesis of atherosclerosis is based on<br />

the premise that factors such as obesity, hypertension, diabetes, and dyslipidemia<br />

share pathways that may play a role both in the preclinical manifestations of<br />

atherosclerosis and the risk of progression to occlusive vascular disease.(Figure<br />

9.2). The clinical factors required for the definition of the metabolic syndrome<br />

are characterized by evidence of inflammation and oxidative stress that is demonstrable<br />

across the spectrum of atherosclerosis and may represent preclinical targets<br />

for risk stratification or therapeutic interventions. 13<br />

CLINICAL MARKERS OF INFLAMMATION<br />

The advent of clinical markers has established a central role for inflammation as<br />

a primary pathogenetic mechanism in the initiation and progression of atherosclerosis.<br />

The concept that atherosclerosis is an inflammatory disease has gained<br />

credence over the past decade. The recognition that obesity is also associated<br />

with a significant inflammatory component has been established from a large<br />

body of clinical evidence that emanated from experimental, epidemiologic,<br />

genetic, and pathologic studies.<br />

Additionally, markers of inflammation have been utilized to link obesity with<br />

other cardiovascular risk factors such as diabetes, hypertension, and dyslipidemia.


144 Oxidative Stress and Inflammatory Mechanisms<br />

Sensitive, reproducible, and specific markers of inflammation are necessary to<br />

establish accurate risk stratification and potentially to monitor response to therapy.<br />

While the bulk of clinical and experimental evidence has centered around the<br />

role of C-reactive protein in the inflammatory state associated with obesity and<br />

atherosclerosis, a number of other pro- and anti-inflammatory proteins have been<br />

employed for clinical stratification including leptin, tumor necrosis factor-α,<br />

adiponectin, interleukins-1, -6, and -10, resistin, fibrinogen, plasminogen activator<br />

inhibitor-1, and others.<br />

C-REACTIVE PROTEIN<br />

Oxidative Stress Inflammation<br />

Endothelial Dysfunction<br />

Atherosclerosis<br />

FIGURE 9.2 The common soil hypothesis.<br />

C-reactive protein is one of the original inflammatory markers that was systematically<br />

analyzed for a potential role in risk stratification for coronary atherosclerosis.<br />

An extensive body of epidemiologic and clinical trial evidence has been<br />

accumulated concerning its role in both vascular disease and obesity. Additionally,<br />

the utility of C-reactive protein as a marker for cardiovascular risk stratification<br />

has continued to expand as improved methods of analysis have been developed.<br />

The original assays for C-reactive protein exhibited wide ranges of normality.<br />

Considerable intra-individual variability was problematic in analyzing the clinical<br />

relevance of relatively minor changes occurring over time.<br />

The subsequent availability of high sensitivity assays has allowed the analysis<br />

of changes of C-reactive protein within the normal range and significantly<br />

improved the clinical utility of this marker in cardiac risk stratification. Multiple<br />

epidemiologic studies in primary prevention have demonstrated that a single<br />

measurement of C-reactive protein is a strong predictor of future vascular events<br />

and is independent of the traditional cardiac risk factors. 14 C-reactive protein<br />

levels are increased in several cardiac risk factors (including truncal obesity)<br />

required by the National Cholesterol Education Program for the diagnosis of the<br />

metabolic syndrome and has added prognostic information to traditional risk<br />

stratification. 15<br />

C-reactive protein has been demonstrated to be an independent predictor of<br />

future cardiac risk and adds prognostic information obtained from traditional lipid


Obesity and Inflammation: Implications for Atherosclerosis 145<br />

screening and the Framingham Risk Score in subjects classified as candidates for<br />

primary prevention. The clinical utility of quantifying C-reactive protein levels<br />

in risk stratification in secondary prevention is more controversial because these<br />

subjects require an aggressive multifaceted therapeutic approach that would<br />

include interventions that lower inflammatory markers. However, C-reactive protein<br />

has been demonstrated to be an independent predictor of future cardiovascular<br />

risk in patients with angiographically documented coronary artery disease and is<br />

also correlated with the number of angiographically determined complex coronary<br />

artery lesions in subjects with acute coronary syndromes. 16 The role that Creactive<br />

protein plays in risk stratification in primary prevention has been clearly<br />

defined over the past decade. However, increasing evidence indicates that it is<br />

also a direct participant in the atherosclerotic process. 17<br />

C-reactive protein is a complex molecule composed of 523 kilodalton subunits<br />

and has been classified as a member of the pentraxin family. The liver is a major<br />

site of the synthesis of C-reactive protein which is at least partially under the<br />

modulation of interleukin-6 produced by adipose cells. 18 C-reactive protein was<br />

determined to be a major component of the immune response system and is<br />

directly involved in the primary migration of inflammatory cells into the subendothelial<br />

space. 19 The recognition that C-reactive protein is an active participant<br />

in the inflammatory process has raised the possibility of a direct role in the<br />

pathogenesis of atherosclerosis.<br />

C-reactive protein has been demonstrated to mediate a variety of pro-inflammatory<br />

and atherosclerotic effects on human aortic endothelial cells following<br />

recognition and binding to FC gamma receptors (CD 32 and CD 64). 20 It has also<br />

been demonstrated to be localized in the vascular intima within regions that<br />

are prone to the development of atherosclerosis. 21 Additionally, C-reactive<br />

protein deposition precedes the appearance of monocytes in the earliest stages of<br />

atherosclerosis.<br />

The monocytes that subsequently give rise to macrophage scavenger cells<br />

have been demonstrated to express a specific receptor for C-reactive protein. The<br />

administration of monoclonal antibodies directed at the receptor eliminates Creactive<br />

protein-mediated monocyte chemotaxis and supports the major role it<br />

plays in the recruitment of inflammatory cells in coronary artery disease. The<br />

localization of C-reactive protein within the subendothelial space is also associated<br />

with the expression of a variety of cellular adhesion molecules, monocyte<br />

chemotactic protein (MCP-1), and plasminogen activator inhibitor (PAI-1). Creactive<br />

protein levels thus appear to be intimately involved with the early phases<br />

of atherosclerosis and are associated with increased risk across the spectrum of<br />

vascular disease.<br />

Additionally, C-reactive protein has been demonstrated to be elevated in<br />

subjects with increased body mass index and obesity which substantiates the role<br />

of inflammation in obesity. The epidemiologic association of elevated levels of<br />

C-reactive protein and cardiovascular risk is robust. 22,23 The presence of elevated<br />

C-reactive protein in obese and insulin-resistant subjects also appears to predict<br />

the subsequent risk of development of type 2 diabetes which also implicates


146 Oxidative Stress and Inflammatory Mechanisms<br />

inflammation as a potential mechanism in the pathogenesis of the metabolic<br />

syndrome. 24<br />

Hygienic measures such as weight loss in obesity have been demonstrated to<br />

lower circulating levels of inflammatory markers such as C-reactive protein by<br />

regulating genes involved with the inflammatory responses in white adipose<br />

tissue. 25 The elevations of C-reactive protein in obesity link the progressive<br />

accumulation of adipose tissue with inflammation and may provide correlation<br />

with cardiovascular risk in obese subjects.<br />

LEPTIN<br />

Leptin was one of the original hormones isolated from adipose tissue and the<br />

discovery subsequently initiated extensive investigations that elucidated its role<br />

in the complex metabolic pathways characterizing obesity. Leptin is directly<br />

synthesized within adipocytes and is a central physiologic factor in the regulation<br />

of both appetite and energy expenditure. It plays a physiologic role in the regulation<br />

of body mass index by increasing energy expenditure and reducing caloric<br />

intake to maintain metabolic balance.<br />

Genetic models such as ob/ob mice that lack the gene encoding for the<br />

synthesis of leptin demonstrate significant increases in body mass index that are<br />

also associated with insulin resistance and high subsequent risks for development<br />

of type II diabetes. 26<br />

Additionally, the administration of leptin in these models has been demonstrated<br />

to result in reductions in both caloric intake and body mass index. The<br />

physiologic action of leptin in the regulation of caloric intake and body mass<br />

requires the presence of normal hormonal levels, receptor function, and signal<br />

transduction. Rare genetic conditions associated with morbid obesity have been<br />

demonstrated to be correlated with either reduced levels of circulating leptin or<br />

associated receptor abnormalities. 27 Leptin has also been utilized as a therapeutic<br />

agent and its administration to subjects with genetically mediated deficiency states<br />

resulted in reductions in body mass index plus improvements in both inflammatory<br />

and metabolic parameters. 28<br />

However, significant differences are present in the forms of obesity prevalent<br />

in the general population compared to subjects with genetically mediated congenital<br />

leptin deficiency. The increased body mass in obese individuals is associated<br />

with a significant capacity to synthesize leptin due to the presence of large<br />

numbers of adipocytes. Leptin resistance is common in subjects with coexistent<br />

insulin resistance, diabetes, and obesity. The precise mechanism that underlies<br />

the high circulating levels of leptin and resistance to the normal physiologic<br />

function of this hormone is complex and multifactorial. The presence of progressively<br />

increasing levels of leptin appears to result in down-regulation of receptor<br />

number or activity, with subsequent resistance to the normal physiologic activity<br />

of the hormone.<br />

The chronic administration of a diet enriched in calories accounted for by<br />

saturated fat had been demonstrated to be associated with a subsequent increase


Obesity and Inflammation: Implications for Atherosclerosis 147<br />

in circulating leptin levels despite a progressive expansion of body mass. The<br />

induced increase in body mass index secondary to caloric excess has been correlated<br />

with a suppression of signaling for leptin due to a progressively increased<br />

expression of SOCS-3 (suppressor of cytokine signaling-3) that significantly<br />

alters the normal physiologic and regulatory effects within hypothalamic centers<br />

in the brain. 29<br />

Leptin also has been demonstrated to have significant metabolic effects in<br />

peripheral tissues that were initially identified in genetic conditions such as<br />

congenital lipodystrophy. Experimental models with this uncommon genetic<br />

condition are associated with minimal adipose tissue deposits in peripheral tissues,<br />

resistance to insulin, and progressive and persistent hyperinsulinemia. The<br />

administration of leptin in subjects with congenital lipodystrophy was demonstrated<br />

to restore the metabolic abnormalities to normal. 30 Leptin is now established<br />

as a major hormone produced by adipocytes and plays a central role in<br />

the regulation of energy homeostasis, insulin sensitivity, and body mass index.<br />

ADIPONECTIN<br />

Adiponectin is an adipocyte-derived protein shown to express beneficial physiologic<br />

effects in clinical conditions associated with the development of coronary<br />

artery disease including significant anti-inflammatory effects 31 (Figure 9.3). It has<br />

been demonstrated to be interrelated with obesity, insulin resistance, endothelial<br />

function and atherosclerosis. Adiponectin may reverse a variety of inflammatory<br />

and metabolic abnormalities which are associated with obesity by increasing the<br />

sensitivity of insulin in the peripheral tissues with resultant improvement in<br />

glucose homeostasis. 32<br />

Adiponectin has also been demonstrated to mediate an improvement in endothelial<br />

function that may be at least partially due to its potent inherent antiinflammatory<br />

effect. Adiponectin has been shown to reduce the synthesis of the<br />

potent tumor necrosis factor-α pro-inflammatory cytokine associated with<br />

endothelial dysfunction. Management of endothelial function associated with<br />

Antiinflammatory<br />

Increased<br />

Angiogenesis<br />

Adiponectin<br />

FIGURE 9.3 Adiponectin and atherosclerosis.<br />

Improved<br />

Insulin<br />

Sensitivity<br />

Improved Endothelial<br />

Function


148 Oxidative Stress and Inflammatory Mechanisms<br />

dyslipidemia or insulin resistance may be enhanced by therapy with drugs such<br />

as PPAR ( peroxisome proliferator-activated receptor) agonists such as fenofibrate<br />

or insulin sensitizers such as thiazolidinediones that also increase adiponectin<br />

levels. 33,34 Circulating adiponectin levels have been evaluated in obese male and<br />

female subjects and proved to be inversely related to body mass index although<br />

the relationship may be more pronounced in women. 35<br />

Reduction in the circulating level of adiponectin is demonstrable in multiple<br />

stages of atherosclerosis ranging from endothelial dysfunction to advanced<br />

obstructive vascular disease. The earliest stage of atherosclerosis is characterized<br />

by inflammatory cell binding to the dysfunctional endothelium which is mediated<br />

by the production of cellular adhesion molecules. The production of adhesion<br />

molecules is mediated by a variety of pro-inflammatory cytokines such as tumor<br />

necrosis factor-α and has been demonstrated to be reduced by adiponectin. The<br />

inverse relationship between adiponectin and tumor necrosis factor-α links the<br />

protective effect of this adipocyte derived protein to the preclinical phases of<br />

coronary artery disease. 36<br />

Adiponectin may also play a protective role in more advanced phases of<br />

vascular obstruction by increasing angiogenesis in chronic ischemia. Experimental<br />

studies have been performed in an adiponectin knock-out mouse model that<br />

employed a chronic hind limb ischemia preparation. The depletion of adiponectin<br />

resulted in a significant impairment of vascular repair of the involved ischemic<br />

area. Adenovirus-mediated supplementation of adiponectin resulted in a significant<br />

improvement in angiogenic repair which implicates a potential role for this<br />

adipocyte-derived protein in the development of collateral flow in chronic<br />

ischemia. 37 Adiponectin thus has significant anti-inflammatory activity and is<br />

reduced in obesity and obstructive vascular disease.<br />

TUMOR NECROSIS FACTOR-α<br />

Tumor necrosis factor-α is a pro-inflammatory cytokine whose levels are<br />

increased in such seemingly disparate clinical conditions as collagen vascular<br />

diseases and congestive heart failure. Modification of the activity of tumor necrosis<br />

factor by receptor blockers has been utilized as a therapeutic target in rheumatoid<br />

arthritis. Tumor necrosis factor-α is directly synthesized in adipocytes<br />

and is felt to play a significant role in the maintenance of both the mass and<br />

metabolism of adipose tissue. 38<br />

The localization of the synthesis of tumor necrosis factor-α within the adipose<br />

tissue also allows for a potential autocrine or paracrine effect that may alter the<br />

numbers and sizes of adipocytes in addition to systemic pro-inflammatory effects.<br />

Tumor necrosis factor-α was found to be associated with enhanced apoptosis of<br />

stem cells that have the potential to develop into mature adipocytes in addition<br />

to fully differentiated adipocytes. 39<br />

The mechanism by which tumor necrosis factor-α induces programmed cell<br />

death is complex and multifactorial. The pro-inflammatory activity of tumor necrosis<br />

factor-α is associated with a variety of proteolytic enzymes that may be involved


Obesity and Inflammation: Implications for Atherosclerosis 149<br />

in progressive and irreversible cellular degradation. 40 Additionally, up-regulation of<br />

the capase gene may also be related to increased levels of tumor necrosis factor-α<br />

and has been demonstrated to be inhibited by the administration of potent antiinflammatory<br />

agents such as glucocorticoids (e.g., dexamethasone). 41<br />

Tumor necrosis factor-α also plays a significant role in lipid metabolism and<br />

the resultant metabolic fate of circulating lipoproteins. The degradation of triglyceride-rich<br />

lipoproteins such as very low density lipoprotein is modulated by<br />

activation of lipoprotein lipase and results in the generation of free fatty acids.<br />

The local availability of free fatty acids within adipose tissue is a significant<br />

determinant of the triglyceride contents of individual adipocytes. Free fatty acids<br />

may be accumulated within adipocytes and subsequently synthesized into<br />

triglycerides under the enzymatic regulation of Acyl-CoA synthetase. Tumor<br />

necrosis factor-α may exert significant effects on fatty acid metabolism and lipid<br />

contents of adipocytes by direct effects on the activity of lipoprotein lipase which<br />

alters substrate availability for triglyceride synthesis.<br />

The intrinsic enzymatic activity of lipoprotein lipase is significantly reduced<br />

by tumor necrosis factor-α with a resultant decrease in the metabolism of triglyceride-rich<br />

lipoproteins. The impaired degradation of triglyceride-rich lipoproteins<br />

decreases the potential for enhanced flux of free fatty acids into adipocytes<br />

and results in the alteration of intracellular lipid concentration. 43<br />

Additionally, tumor necrosis factor-α is associated with direct down-regulation<br />

of acyl Co-A synthetase which may also contribute to the degree of triglyceride<br />

synthesis and lipid accumulation within adipocytes. 43<br />

The role of tumor necrosis factor-α in the manifestations of inflammation in<br />

obesity in human subjects is complex. The local synthesis of tumor necrosis<br />

factor-α in the adipocytes of obese subjects is increased and sustained weight<br />

loss has been demonstrated to reduce circulating levels. 44 However, conflicting<br />

results have been reported concerning the relationship of circulating tumor necrosis<br />

factor-α and body mass index that may be secondary to methodologic problems.<br />

Additionally, genetic models for obesity have demonstrated that increased<br />

circulating levels of tumor necrosis factor-α are associated only with extreme<br />

increases in body weight. 45<br />

Obesity and increased body mass index are frequently associated with the<br />

development of insulin resistance and adult onset diabetes. Inflammation has been<br />

postulated to play a significant role in the pathogenesis of insulin resistance in<br />

obese subjects. The role of tumor necrosis factor-α in the pathogenesis of insulin<br />

resistance has been evaluated in experimental studies. Obese subjects with associated<br />

insulin resistance showed increased circulating levels of tumor necrosis<br />

factor-α that were compatible with the possibility of a cause-and-effect relationship.<br />

Tumor necrosis factor-α may induce insulin resistance by several mechanisms<br />

including down-regulation of the insulin receptor, decreased GLUT 4<br />

synthesis, and altered lipolytic activity. 46 The interrelationship between obesity<br />

and tumor necrosis factor is complex and a hypothesis that explains all aspects<br />

remains elusive due to multiple genetic, gender, and metabolic issues. However,


150 Oxidative Stress and Inflammatory Mechanisms<br />

it is clear that tumor necrosis factor-α plays a central role in adipocyte function<br />

and mass and may have implications for insulin resistance and diabetes.<br />

HEMOSTATIC PARAMETERS<br />

Obesity has been correlated with a variety of abnormalities of the hemostatic<br />

system including alterations in the levels of fibrinogen, von Willebrand factor,<br />

and plasminogen activator inhibitor. Fibrinogen, which is primarily synthesized<br />

in the liver, is intimately related to the coagulation cascade and additionally is<br />

an acute phase reactant. Fibrinogen is thus closely linked to inflammation and is<br />

demonstrated to be increased following a variety of inflammatory stimuli.<br />

Obesity has been demonstrated to be associated with increased circulating<br />

levels of fibrinogen. 47 However, the relationship between fibrinogen levels and<br />

body mass index or localization (truncal versus peripheral) of adipose tissue is<br />

controversial. The metabolic syndrome (for which truncal obesity is one of the<br />

diagnostic criteria) has been correlated with abnormalities in multiple coagulation<br />

parameters. Fibrinogen levels are increased in individuals fulfilling criteria for a<br />

diagnosis of metabolic syndrome which may contribute to the increased risk of<br />

hypercoagulability associated with the classical risk factors such as hypertension,<br />

diabetes, and dyslipidemia. 48<br />

Endothelial dysfunction is associated with a variety of abnormalities in the<br />

coagulation and fibrinolytic systems. The von Willebrand factor is a large glycoprotein<br />

that is synthesized at least partially in endothelial cells and may represent<br />

a clinical marker for endothelial dysfunction. However, the relationship between<br />

von Willebrand factor and obesity is complex and difficult to elucidate because<br />

the glycoprotein is not directly synthesized by adipocytes. However, the levels<br />

of von Willebrand factor antigen have been significantly correlated with the degree<br />

of visceral fat in subjects who are either overweight or definitely obese. 49<br />

Obesity is frequently associated with alterations of the balance between<br />

thrombotic and fibrinolytic parameters. The primary inhibitor of tissue plasminogen<br />

activation is mediated by plasminogen activator inhibitor-1 (PAI-1) which<br />

decreases the rate of breakdown of fibrin clots (Figure 9.4). Risk factors such as<br />

obesity, hypertension, hypertriglyceridemia, and diabetes are frequently characterized<br />

by elevations of circulating levels of plasminogen activator inhibitor. The<br />

presence of a relative excess of circulating plasminogen activator inhibitor is<br />

associated with an increased risk of occlusive intravascular thrombus due to<br />

impaired fibrinolysis. Plasminogen activator inhibitor is synthesized in response<br />

to inflammation by endothelial cells, adipocytes, hepatocytes, and platelets. 50<br />

Additionally, the levels of plasminogen activator inhibitor have been correlated<br />

with both obesity and increased body mass index which establishes an association<br />

of obesity, hypercoagulability, and inflammation. 51<br />

Abnormalities of plasminogen activator inhibitor levels are further correlated<br />

with the degree of visceral fat, which emphasizes the role of the distribution of<br />

adipose tissue in cardiovascular risk. 52 The relationship between visceral fat<br />

accumulation and abnormalities in the fibrinolytic pathway is emphasized by the


Obesity and Inflammation: Implications for Atherosclerosis 151<br />

fact that reductions in plasminogen activator inhibitor are best correlated with a<br />

decrease in the amount of visceral adipose rather than body weight per se. 53<br />

The current definitions of the metabolic syndrome do not include components<br />

of the hemostatic or fibrinolytic system. However, the risk factors (e.g., diabetes,<br />

hypertension, obesity, and dyslipidemia) included in the criteria for diagnosis<br />

have inflammatory bases and are also frequently associated with either hypercoagulability<br />

or impaired fibrinolysis. 54 Markers of inflammation or fibrinolysis may<br />

be included in the criteria for metabolic syndrome as further studies evolve and<br />

more sensitive and specific indicators of cardiac risk are evaluated.<br />

RESISTIN<br />

Plasminogen<br />

Plasminogen<br />

T-PA<br />

T-PA<br />

PAI-I<br />

Plasmin<br />

TG<br />

All<br />

Lp(a)<br />

TNF<br />

Insulin<br />

Plasmin<br />

FIGURE 9.4 Endothelial dysfunction and fibrinolysis.<br />

Fibrinolysis<br />

Thrombosis<br />

Resistin is a recently described pro-inflammatory cytokine that is produced within<br />

adipose tissue and has been linked to both obesity and the risk for development<br />

of diabetes mellitus. 55 However, the role of resistin in the characteristic obesity<br />

of humans is controversial despite the compelling data from animal models.<br />

Resistin is a cysteine-rich polypeptide that is directly synthesized by adipocytes. 56<br />

Circulating levels of resistin were shown to be increased in experimental<br />

models that demonstrated impaired insulin sensitivity and diabetes. Additionally,<br />

circulating resistin levels were reduced by insulin sensitizing agents such as<br />

thiazolidinediones. 57 The increased levels of resistin in obesity have been demonstrated<br />

to be generated by direct synthesis from activated macrophages that are<br />

localized within adipose tissue and regulated by PPAR gamma activators. 58<br />

The glucose intolerance associated with obesity may also be linked to<br />

increased levels of resistin via complex metabolic and inflammatory pathways.<br />

Experimental studies utilizing transgenic mice with high circulating concentrations<br />

of resistin in the presence of normal weight showed increased glucose<br />

production in the setting of a hyperinsulinemic–euglycemic clamp. 59 The fact that<br />

elevated levels of resistin are associated with impaired glucose homeostasis is at<br />

least partially explained by the associated increased expression of the phosphoenolpyruvate<br />

carboxykinase hepatic enzyme that serves as a key enzyme in


152 Oxidative Stress and Inflammatory Mechanisms<br />

glucose metabolism. The relationship between resistin and inflammation has been<br />

demonstrated by an association with inflammatory mediators such as tumor<br />

necrosis factor-α and C-reactive protein. 60,61 Resistin has been hypothesized to<br />

play multiple roles in obesity and diabetes secondary to the induction of inflammation<br />

and a secondary alteration of the metabolism of glucose due to the<br />

associated insulin resistance.<br />

OTHER INFLAMMATORY MARKERS<br />

A variety of other markers have been proposed as clinically relevant parameters<br />

that have utility in the identification of inflammation and potential cardiovascular<br />

risk stratification. Interleukin-6, interleukin-10, interleukin-1, adhesion molecules,<br />

and visfatin have all been identified as potentially valuable clinical markers.<br />

However, the bulk of clinical and experimental data related to obesity, diabetes,<br />

and inflammation were derived from the identification of C-reactive protein and<br />

improved high sensitivity measurements that have allowed accumulation of a<br />

large body of clinical and experimental evidence that substantiates the role of<br />

inflammation in obesity<br />

ATHEROSCLEROSIS <strong>AND</strong> INFLAMMATION<br />

The concept of atherosclerosis as an inflammatory disorder is supported by an<br />

ever-increasing body of pathologic, experimental, epidemiologic, and clinical<br />

evidence. 62 The delineation of the role that inflammation plays as a primary factor<br />

in the initiation and progression of vascular disease has led to enhanced understanding<br />

of the basic biologic properties of atherosclerotic plaques, risk stratification,<br />

and therapeutic strategies.<br />

The cellular elements involved in inflammation can be demonstrated to be<br />

present to variable degrees in multiple stages in the atherosclerotic process.<br />

Additionally, inflammation also can be related to the classic cardiovascular risk<br />

factors including obesity. The process by which inflammatory cells are localized<br />

into the subendothelial space and contribute to atherosclerosis is complex and<br />

requires cellular recognition, binding, migration, and transformation. The normal<br />

endothelium is not associated with significant adherence of inflammatory cells.<br />

However, the dysfunctional endothelium expresses a variety of adhesion molecules<br />

such as vascular cell adhesion molecule-1 (VCAM-1) and members of the<br />

selectin (E and P) family. 63<br />

The progressive expression of adhesion molecules on the dysfunctional endothelium<br />

provides the means for the recognition, binding, tethering, and transmigration<br />

of circulating inflammatory cells. The accumulation of inflammatory cells<br />

is an initial phase of atherosclerosis. Additionally, cytokines and chemokines are<br />

small proteins that are intimately involved in the transmigration and concentration<br />

of inflammatory cells into vascular regions prone to develop atherosclerosis. 64<br />

Cytokines are involved in enhanced migration of several inflammatory cell lines<br />

of which the monocyte is the prototype example.


Obesity and Inflammation: Implications for Atherosclerosis 153<br />

Monocyte chemoattractant protein (MCP-1) is a primary regulator of the<br />

accumulation and concentration of monocytes into the subendothelial spaces in<br />

the initial stages of atherosclerosis. Monocytes play a significant role in the<br />

cellular inflammatory response and also can transform into scavenger cells that<br />

are involved in the recognition and binding of oxidized low density lipoprotein<br />

with subsequent generation of lipid-laden foam cells. The capacity of the monocyte<br />

to transform into a macrophage scavenger cell line is under the regulation<br />

of a variety of colony stimulating factors. 65 The monocyte scavenger receptor<br />

recognizes both native and modified low-density lipoprotein. However, the binding<br />

and uptake of non-modified low-density lipoprotein is quantitatively minimal.<br />

In contrast, low-density lipoproteins modified by processes such as oxidation<br />

or glycation are progressively internalized at a relatively rapid rate by macrophages<br />

and are associated with progressive lipid accumulation and the subsequent<br />

development of atherosclerosis. In contrast to the classic LDL receptor, the<br />

number or activity of the scavenger receptor is not down-regulated by progressive<br />

accumulation of intracellular lipid.<br />

The macrophage that originates from the inflammatory cell line also plays a<br />

significant role in the progression and ultimate fate of the atherosclerotic plaque.<br />

The monocyte–macrophage cellular elements are gradually depleted due either<br />

to apoptosis or the cytotoxic effects of a variety of noxious stimuli. The degradation<br />

of cellular constituents results in the release of lipids and proteinaceous<br />

debris into the plaque and thus has implications for vulnerability and rupture.<br />

Macrophages also play a significant role in the local degree of oxidative stress<br />

demonstrated to be secondary to the capacity to synthesize cytotoxic reactive<br />

oxygen species. 66 The concentration and rate of production of reactive oxidative<br />

species are major contributors to the initiation and progression of atherosclerosis<br />

effects on lipid oxidation and cellular damage.<br />

The degree of inflammation also plays a significant role in the vulnerability<br />

of the atherosclerotic plaque and subsequent risk for rupture. High-grade inflammation<br />

is associated with a progressive reduction in plaque stability and resultant<br />

increased vulnerability. Macrophages have the capacity to elaborate a variety of<br />

proteolytic enzymes that may be involved in the net balance between intraplaque<br />

matrix synthesis and degradation. Collagenase and gelatinase are matrix metalloproteinases<br />

that demonstrate the capacity to progressively degrade the stabilizing<br />

matrix protein constituents and result in a progressively increased risk for<br />

rupture due to reduction in the tensile strength of the plaque. 67<br />

Increased plaque vulnerability is characterized by an enhanced risk of erosion<br />

or frank rupture of the fibrous cap. Destruction of the structural integrity of the<br />

fibrous cap results in exposure of platelets and coagulation factors to the highly<br />

thrombogenic lipid-rich core localized within the atherosclerotic plaque. Procoagulants<br />

such as tissue factor are produced within the lipid-laden foam cell<br />

and are associated with increased risk for vascular occlusion due to enhanced<br />

synthesis of activated clotting factors VII and X. 68 Macrophages are also involved<br />

in the synthesis of inhibitors to plasminogen activators (e.g., PAI-I), resulting in<br />

a decreased capacity to lyse a potentially occlusive intravascular thrombus. 69


154 Oxidative Stress and Inflammatory Mechanisms<br />

Inflammation is thus a major contributor to all phases of atherosclerosis from<br />

endothelial dysfunction to acute myocardial infarction. The major cardiac risk<br />

factors such as diabetes, hypertension, and dyslipidemia are all associated with<br />

a degree of inflammation and oxidative stress. Recent evidence has centered<br />

around the role of adipocytes in inflammation and a large body of work relates<br />

obesity to the inflammatory response.<br />

OBESITY <strong>AND</strong> INFLAMMATION<br />

The major cardiac risk factors frequently coexist and share multiple common<br />

metabolic pathways. Diabetes, hypertension, the use of tobacco products, and<br />

dyslipidemia have all been demonstrated to be associated with inflammation.<br />

However, the concept that obesity is characterized by active inflammation is a<br />

relatively new concept that initially was not widely accepted. The previous perception<br />

of obesity was that adipose tissue simply represented a passive storage<br />

depot for excess energy resulting from increased caloric intake relative to energy<br />

expenditure.<br />

Adipose tissue was subsequently demonstrated to be a metabolically active<br />

endocrine organ and the resultant increase in body mass index was implicated in<br />

the pathogenesis of diabetes, hypertension, and dyslipidemia. Adipose tissue has<br />

been shown to be intimately involved in a variety of inflammatory processes and<br />

the mechanisms have been partially clarified. Adipose tissue is the source of a<br />

number of cytokines that regulate the production of inflammatory mediators<br />

including C-reactive protein, tumor necrosis factor-α, interleukin-6, etc.<br />

Insulin resistance is also thought to play a significant role in the pathogenesis<br />

of several cardiovascular risk factors in obese subjects. Individuals with elevated<br />

circulating levels of insulin due to impaired peripheral resistance have increased<br />

free fatty acid release from adipocytes and enhanced delivery to the liver. The<br />

increased bioavailability of free fatty acids as a synthetic substrate results in<br />

enhanced hepatic production of very low-density lipoprotein. Additionally, the<br />

enzymatic activity of lipoprotein lipase is decreased in obese subjects with insulin<br />

resistance. The reduced catabolic rates in combination with hepatic overproduction<br />

of triglyceride-rich particles results in an atherogenic dyslipidemic phenotype<br />

characterized as the lipid triad (hypertriglyceridemia, low HDL, and small dense<br />

LDL) (Figure 9.5 and 9.6).<br />

Additionally, hyperinsulinemia is associated with vascular smooth hypertrophy,<br />

enhanced sympathetic activity, and increased renal sodium reabsorption with<br />

the resultant development of systemic hypertension. 70 Adipocytes have been<br />

implicated as the initial sites of inflammation in obesity and predate the involvement<br />

of other organ systems. C-reactive protein has been demonstrated to modulate<br />

the movement of inflammatory cells into adipose tissue via the production<br />

of a variety of chemokines. Monocyte migration and subsequent transformation<br />

into macrophages within adipose tissue are key factors in the self-perpetuating<br />

low-grade inflammation associated with obesity.


Obesity and Inflammation: Implications for Atherosclerosis 155<br />

Relative Insulin Deficiency<br />

↓ Lipoprotein Lipase ↑ Free Fatty Acids<br />

↓ Clearing of VLDL<br />

FIGURE 9.5 Insulin deficiency and dyslipidemia.<br />

Increased<br />

Sympathetic<br />

Tone<br />

↑ VLDL ↓ HDL<br />

Small Dense LDL<br />

Insulin Resistance<br />

Hyperinsulinemia<br />

Renal Sodium<br />

Reabsorption<br />

Hypertension<br />

FIGURE 9.6 Insulin levels and hypertension.<br />

↑ Hepatic Synthesis<br />

of VLDL<br />

Vascular<br />

Hypertrophy<br />

The demonstration that adipocytes and monocytes share pathways in the<br />

immune response system emphasizes the role that adipose tissue plays in the<br />

interaction between inflammation and atherosclerosis. Additionally, the cellular<br />

precursors to mature adipocytes were demonstrated to have the ability, similar to<br />

the monocyte–macrophage system, to act as phagocytes. The adipocytes localized<br />

within fat depots may thus participate in a primary manner in the inflammatory<br />

response, systemic hypertension, and atherosclerosis.<br />

The inflammatory response within adipose tissue may at least be partially<br />

expressed under genetic control. Experimental models that utilized animals with<br />

genetically transmitted obesity allowed the examination of factors that control<br />

and modulate inflammatory states associated with increases in body mass. Multiple<br />

genes have been implicated in the inflammatory response associated with<br />

obesity. Genes that regulate the monocyte–macrophage system have been demonstrated<br />

to be present in white adipose tissue. The experimental studies in the<br />

obese mouse models have demonstrated the expression of a large number of<br />

transcripts that can be correlated with body mass.


156 Oxidative Stress and Inflammatory Mechanisms<br />

Biochemical markers for macrophages were significantly correlated with both<br />

body mass and the size of the adipocytes. Analysis of the genes that had the<br />

highest correlations with obesity demonstrated that approximately 30% were<br />

involved in encoding for proteins that were characteristic of macrophages. 71<br />

Tumor necrosis factor-α is a pro-inflammatory cytokine demonstrated to be<br />

produced by macrophages and is localized within adipose tissue. The potential<br />

thus exists for a self-perpetuating cycle of increased macrophage infiltration<br />

coupled with enhanced synthesis of cytokines within adipose tissue and a resultant<br />

persistent inflammatory milieu.<br />

The establishment of an inflammatory environment within adipose tissue may<br />

also play a significant role in the pathogenesis of insulin resistance. However,<br />

significant differences may exist among genetically mediated obesity, insulin<br />

resistance, and an increased body mass that is purely secondary to excessive<br />

dietary intake of calories. Chronic inflammation has been linked to the pathogenesis<br />

of both type 2 diabetes and insulin resistance. 72 Insulin resistance is due to<br />

a failure of circulating hormonal levels to mediate the normal metabolic handling<br />

of glucose despite progressively increased concentrations and has been related<br />

to the presence of inflammation. However, the underlying pathogenesis and interrelationship<br />

of the various metabolic pathways is not totally clear.<br />

Inflammation and macrophage-specific genes have been demonstrated to play<br />

direct roles in both models of genetic obesity and the obesity associated with a<br />

high fat diet. 73 Macrophages and inflammatory-specific genes are significantly<br />

up-regulated in the white adipose tissues of both diet-induced obesity and genetic<br />

mouse models. The up-regulated genes precede a significant increase in the level<br />

of circulating insulin. Histologic examination demonstrates that macrophage infiltration<br />

is present in the white adipose tissue and is also associated with evidence<br />

of cellular destruction of adipocytes.<br />

Therapy with rosiglitazone, an insulin sensitizing agent, was associated with<br />

a significant down-regulation of the macrophage-originated genes. Thus, inflammation<br />

and macrophage infiltration may play significant roles in both diet-induced<br />

obesity and genetically induced obesity and result in progressive insulin resistance<br />

due to a chronic inflammatory and cytotoxic response initiated in adipose tissue.<br />

MODIFICATION OF INFLAMMATION IN OBESITY<br />

The identification of obesity as a condition associated with a definite inflammatory<br />

component has significant clinical implications relative to targets of therapy (i.e.,<br />

modification of body mass index in and of itself, improving the function of<br />

adipose tissue, and normalizing the levels of inflammatory markers). Therapeutic<br />

interventions that alter the inflammatory response may also be associated with<br />

restoration of the normal physiologic function of adipose tissue.<br />

The initial and primary foci of therapy in obesity involve lifestyle modification<br />

with restriction of calories and saturated fat coupled with increased energy<br />

expenditure through physical activity. Increased physical activity has been demonstrated<br />

to decrease body mass index, improve dysfunctional adipose tissue, and


Obesity and Inflammation: Implications for Atherosclerosis 157<br />

reduce levels of inflammatory markers. However, subjects who are obese but<br />

metabolically healthy may require alternative approaches to improve long-term<br />

cardiovascular risk. As risk factor stratification proceeds beyond tabulation of the<br />

classical risk factors, modification of inflammatory markers may possibly become<br />

a primary target for therapy with pharmacologic or lifestyle interventions.<br />

The role of modification of inflammatory markers as targets for therapy has<br />

been addressed in a variety of clinical trials utilizing lifestyle or pharmacologic<br />

interventions.<br />

C-REACTIVE PROTEIN<br />

C-reactive protein is a major inflammatory marker and a number of lifestyle or<br />

pharmacologic interventions have been demonstrated to reduce circulating levels<br />

of this marker. The effect of weight loss on C-reactive protein has been prospectively<br />

determined in controlled clinical trials utilizing a variety of dietary interventions.<br />

The role of implementation of a Mediterranean style diet low in calories<br />

but associated with an increase in monounsaturated fat coupled with an increase<br />

in physical activity and designed to decrease body weight by 10% has been<br />

analyzed.<br />

Body mass index, interleukin-6, and C-reactive protein were all significantly<br />

reduced by the combination of diet and lifestyle intervention. Additionally, adiponectin<br />

was increased, thus demonstrating that diet-induced weight loss and<br />

enhanced physical activity improved the balance of pro- and anti-inflammatory<br />

mediators. 74 Risk factors that compromise the components of the metabolic syndrome<br />

have inflammation as a common denominator.<br />

The impact of the institution of a Mediterranean style diet on endothelial<br />

function and markers of inflammation in subjects who fulfill the criteria for the<br />

diagnosis of the metabolic syndrome has been investigated. The dietary intervention<br />

consisted of an increase in the consumption of vegetables, whole grains,<br />

fruits, olive oil, and nuts. The Mediterranean diet resulted in significant reductions<br />

of C-reactive protein and improvements in insulin resistance. Endothelial function<br />

score as analyzed by blood pressure and platelet aggregation responses to larginine<br />

was also improved by dietary therapy. The components of the metabolic<br />

syndrome were normalized to a greater degree in subjects who were randomized<br />

to receive intensive dietary interventions. 75<br />

The effect of statin therapy on inflammatory markers had been presumed to<br />

be secondary to lipid lowering. However, statins have been clearly demonstrated<br />

to possess a variety of pleiotropic or non-lipid effects. Statins may play a significant<br />

role in leukocyte trafficking and T cell activation by binding to a novel<br />

allosteric site within the β-2 integrin leukocyte function associated antigen-1. 76<br />

Statin therapy has also been shown to reduce C-reactive protein levels and has<br />

been correlated with clinical outcomes in prospective clinical studies (albeit in<br />

post-hoc analyses).<br />

The role of statin therapy in lipid lowering and alteration of inflammatory<br />

markers was initially analyzed in the Cholesterol and Recurrent Events (CARE)


158 Oxidative Stress and Inflammatory Mechanisms<br />

trial that prospectively compared pravastatin to placebo in subjects with relatively<br />

normal cholesterol levels who had suffered acute myocardial infarctions. 77 The<br />

highest cardiovascular risks occurred in subjects who had significant evidence of<br />

inflammation and were randomized to receive placebo. Pravastatin therapy<br />

reduced markers of inflammation (C-reactive protein and serum amyloid A) in<br />

addition to beneficial effects on lipid profiles.<br />

The utilization of C-reactive protein in predicting response to statin therapy<br />

in the primary prevention of atherosclerosis was the subject of the Air Force/Texas<br />

Coronary Atherosclerosis Prevention Study (AFCAPS/TEXCAPS) analyzing the<br />

effect of lovastatin in 5742 relatively low risk subjects over 5 years. 78 The effect<br />

of lovastatin was analyzed after stratification of subjects by the ratio of plasma<br />

cholesterol:HDL and levels of C-reactive protein. Lovastatin reduced coronary<br />

events irrespective of C-reactive protein levels in subjects with elevated lipid<br />

ratios. However, it was also effective in subjects whose lipid ratios were below<br />

the median but associated with elevated C-reactive protein levels, inferring benefits<br />

of statin therapy in subjects with evidence of inflammation despite the<br />

presence of normal lipid levels.<br />

C-reactive protein is associated with an increased density of the AT-1 angiotensin<br />

receptor, and modulators of the renin angiotensin system may be associated<br />

with an anti-inflammatory effect. Losartan, a renin angiotensin receptor blocker,<br />

has been associated with reductions in levels of C-reactive protein although this<br />

may not be a class effect. 79,80<br />

Diabetes and insulin resistance have been correlated with a variety of markers<br />

of inflammation including C-reactive protein that additionally may play a role in<br />

pathogenesis. 81 The treatment of insulin resistance and diabetes with insulin<br />

sensitizing agents such as rosiglitazone may demonstrate beneficial affects on<br />

both circulating glucose levels and markers of inflammation such as C-reactive<br />

protein. 82 Statins have been demonstrated to reduce the morbidity and mortality<br />

in diabetes in multiple large scale trials and are also effective even if baseline<br />

lipids are relatively normal. 83,84 Additionally, HMG CoA reductase inhibitor therapy<br />

has been demonstrated to reduce the incidence of diabetes in subjects in<br />

primary prevention, implicating a possible non-lipid effect of statin therapy. 85 The<br />

multivariate predictors of the development of diabetes in the West of Scotland<br />

Coronary Prevention Study were glucose, body mass index, and triglyceride<br />

levels. Interestingly, randomization to pravastatin therapy resulted in a 30% reduction<br />

in the incidence of diabetes. The precise mechanism involved in the reduction<br />

of the incidence of diabetes could not be delineated but a potential benefit from<br />

an anti-inflammatory effect was postulated.<br />

Additionally, angiotensin-converting enzyme inhibition has demonstrated<br />

clinical benefits across the spectrum of atherosclerosis in multiple clinical trials.<br />

Angiotensin II has been shown to demonstrate a significant inflammatory effect<br />

and the use of ramipril was demonstrated to decrease the incidence of diabetes<br />

in the Heart Outcomes Prevention Evaluation (HOPE) study. 86 Additionally, in<br />

an extension of the original HOPE trial that continued 30 months after the


Obesity and Inflammation: Implications for Atherosclerosis 159<br />

termination of the initial trial, ramipril use was shown to be associated with<br />

sustained reductions in the incidence of new onset diabetes.<br />

TUMOR NECROSIS FACTOR-α<br />

Tumor necrosis factor-α is well established as a pro-inflammatory cytokine<br />

although the effects of interventions to modify circulating levels in obesity or<br />

cardiovascular disease lack the broad data base associated with C-reactive protein.<br />

However, tumor necrosis factor-α is directly synthesized in adipose tissue and<br />

may be a target for therapy in obesity.<br />

Clinical studies focused on the role of weight reduction in obese subjects as<br />

a means to reduce insulin resistance, utilizing hypocaloric diets and increased<br />

physical activity, have demonstrated decreases in circulating levels of adipokines<br />

including tumor necrosis factor-α. 87 However, the individual studies were small<br />

and the reduction in tumor necrosis factor approached but did not reach statistical<br />

significance.<br />

Hypolipidemic therapy with statins has been demonstrated to reduce tumor<br />

necrosis factor-α levels. Interestingly, the reduction in levels also correlated with<br />

a significant reduction in matrix metalloproteinase levels, which provides a link<br />

among statin therapy, inflammation, and plaque stability. 88 The increase in tumor<br />

necrosis factor-α levels associated with obesity and insulin resistance may be<br />

successfully managed by treatment with PPAR gamma agonists.<br />

The anti-inflammatory effect of rosiglitazone has been evaluated in both<br />

diabetic and non-diabetic obese subjects. Tumor necrosis factor-α levels were<br />

reduced in obese subjects compatible with significant anti-inflammatory effects<br />

of the PPAR gamma agonists. Tumor necrosis factor has been established as a<br />

valuable marker of inflammation and may be reduced by a variety of lifestyle<br />

and pharmacologic interventions that improve lipids, insulin sensitivity, and blood<br />

pressure in addition to potential anti-inflammatory effects. However, further clinical<br />

trials are necessary to demonstrate the clinical effects of modifying circulating<br />

levels of tumor necrosis factor-α.<br />

HEMOSTATIC FACTORS<br />

Hemostatic factors such as fibrinogen and plasminogen activator inhibitor are<br />

increased in obesity and are also linked to inflammation. However, lifestyle and<br />

pharmacologic interventions produced variable effects on these parameters and<br />

a consensus view of beneficial interventions has not been established due to<br />

conflicting results. However, in some studies, statin therapy was demonstrated to<br />

improve hemostatic parameters.<br />

Atorvastatin has been demonstrated to improve global fibrinolytic activity<br />

and reduce plasminogen activator inhibitor coupled with an increase in tissue<br />

plasminogen activator levels in hypercholesterolemic patients. 89 The effect of<br />

simvastatin on hemostatic and fibrinolytic factors has been studied in diabetic<br />

patients. In addition to the beneficial effects on total and low-density lipoprotein


160 Oxidative Stress and Inflammatory Mechanisms<br />

cholesterol levels, simvastatin was demonstrated to reduce circulating plasminogen<br />

activator inhibitor concentration and prothrombinase activity. 90<br />

Additionally, lifestyle interventions have been demonstrated to beneficially<br />

alter fibrinolytic activity in some studies of obese patients. Reduction in weight<br />

and increased physical activity in obese subjects with insulin resistance have been<br />

demonstrated to result in improvement in levels of plasminogen activator inhibitor.<br />

91 Hemostatic factors may be related to an imbalance in fibrinolytic activity<br />

and impaired capacity to lyse intravascular thrombi. While hemostatic factors<br />

represent attractive targets in obesity due to their relationship with inflammation<br />

and may be modified by pharmacologic interventions or lifestyle changes, a<br />

consensus view regarding their role as a modifiable risk factor remains controversial<br />

and will require larger prospective studies.<br />

SUMMARY<br />

The incidence and prevalence of obesity are markedly increasing in the industrialized<br />

world and have reached epidemic proportions. Obesity had been regarded<br />

as a passive depot for excess energy. Recent studies have demonstrated that<br />

adipose tissue is a dynamic and metabolically active organ system with multiple<br />

endocrine functions. Obesity is significantly interrelated with a variety of classical<br />

cardiac risk factors including diabetes, hypertension, dyslipidemia, and hemostatic<br />

abnormalities. Adipose tissue has been demonstrated to produce a number<br />

of pro-inflammatory cytokines that establish a self perpetuating low-grade inflammatory<br />

state. Lifestyle modifications such as caloric restriction and increased<br />

physical activity have been demonstrated to reduce both body weight and a variety<br />

of inflammatory markers.<br />

Pharmacologic interventions such as statin therapy, modulators of the renin<br />

angiotensin system, and insulin sensitizers have been demonstrated to exhibit not<br />

only effects on blood pressure, lipids and glucose, but they also significantly<br />

reduce inflammatory mediators. Thus obesity should be viewed as a highly active<br />

metabolic state with strong components of inflammation and oxidative stress that<br />

predispose a patient to coronary artery disease and should represent an active<br />

target for interventions designed to reduce risks of coronary artery disease.<br />

REFERENCES<br />

1. Mokdad, A.H. et al. The continuing increase of diabetes in the U.S. Diabetes Care<br />

24, 412 (2001).<br />

2. Ogden, C.L. et al. Prevalence of overweight and obesity in the United States,<br />

1999–2004. JAMA 295, 1549 (2006).<br />

3. National Task Force on the Prevention and Treatment of Obesity. Long-term<br />

pharmacotherapy in the management of obesity. JAMA 276, 1907 (1996).<br />

4. Stevens, J. et al. The effect of age on the association between body-mass index<br />

and mortality. New Engl J Med 338, 1 (1998).


Obesity and Inflammation: Implications for Atherosclerosis 161<br />

5. Fontaine, K.R., Redden, D.T., Wang, C., Westfall, A.O., and Allison, D.B. Years<br />

of life lost due to obesity. JAMA 289, 187 (2003).<br />

6. Hu, F.B. et al. Adiposity as compared with physical activity in predicting mortality<br />

among women. New Engl J Med 351, 2694 (2004).<br />

7. Wei, M. et al. Relationship between low cardiorespiratory fitness and mortality<br />

in normal-weight, overweight, and obese men. JAMA 282, 1547 (1999).<br />

8. Lee, C.D., Blair, S.N., and Jackson, A.S. Cardiorespiratory fitness, body composition,<br />

and all-cause and cardiovascular disease mortality in men. Am J Clin Nutr<br />

69, 373 (1999).<br />

9. Wessel, T.R. et al. Relationship of physical fitness vs body mass index with<br />

coronary artery disease and cardiovascular events in women. JAMA 292, 1179<br />

(2004).<br />

10. Weinstein, A.R. et al. Relationship of physical activity vs body mass index with<br />

type 2 diabetes in women. JAMA 292, 1188 (2004).<br />

11. Grundy, S.M. et al. Diagnosis and management of the metabolic syndrome: an<br />

American Heart Association/National Heart, Lung, and Blood Institute Scientific<br />

statement. Circulation 112, 2735 (2005).<br />

12. Kahn, R., Buse, J., Ferrannini, E., and Stern, M. The metabolic syndrome: time<br />

for a critical appraisal: joint statement from the American Diabetes Association<br />

and the European Association for the Study of Diabetes. Diabetes Care 28, 2289<br />

(2005).<br />

13. Stern, M.P. Diabetes and cardiovascular disease: the “common soil” hypothesis.<br />

Diabetes 44, 369 (1995).<br />

14. Ridker, P.M. Clinical application of C-reactive protein for cardiovascular disease<br />

detection and prevention. Circulation 107, 363 (2003).<br />

15. Ridker, P.M., Buring, J.E., Cook, N.R., and Rifai, N. C-reactive protein, the<br />

metabolic syndrome, and risk of incident cardiovascular events: an 8-year followup<br />

of 14,719 initially healthy American women. Circulation 107, 391 (2003).<br />

16. Arroyo-Espliguero, R. et al. C-reactive protein elevation and disease activity in<br />

patients with coronary artery disease. Eur Heart J 25, 401 (2004).<br />

17. Jialal, I., Devaraj, S., and Venugopal, S.K. C-reactive protein: risk marker or<br />

mediator in atherothrombosis? Hypertension 44, 6 (2004).<br />

18. Moshage, H.J. et al. The effect of interleukin-1, interleukin-6 and its interrelationship<br />

on the synthesis of serum amyloid A and C-reactive protein in primary<br />

cultures of adult human hepatocytes. Biochem Biophys Res Commun 155, 112<br />

(1988).<br />

19. Han, K.H. et al. C-reactive protein promotes monocyte chemoattractant protein-<br />

1-mediated chemotaxis through upregulating CC chemokine receptor 2 expression<br />

in human monocytes. Circulation 109, 2566 (2004).<br />

20. Devaraj, S., Du Clos, T.W., and Jialal, I. Binding and internalization of C-reactive<br />

protein by Fc gamma receptors on human aortic endothelial cells mediates biological<br />

effects. Arterioscler Thromb Vasc Biol 25, 1359 (2005).<br />

21. Torzewski, M. et al. C-reactive protein in the arterial intima: role of C-reactive<br />

protein receptor-dependent monocyte recruitment in atherogenesis. Arterioscler<br />

Thromb Vasc Biol 20, 2094 (2000).<br />

22. Jialal, I. and Devaraj, S. Inflammation and atherosclerosis: the value of the highsensitivity<br />

C-reactive protein assay as a risk marker. Am J Clin Pathol 116, Suppl<br />

S108 (2001).


162 Oxidative Stress and Inflammatory Mechanisms<br />

23. Rifai, N. and Ridker, P.M. High-sensitivity C-reactive protein: a novel and promising<br />

marker of coronary heart disease. Clin Chem 47, 403 (2001).<br />

24. Muntner, P., He, J., Chen, J., Fonseca, V., and Whelton, P.K. Prevalence of nontraditional<br />

cardiovascular disease risk factors among persons with impaired fasting<br />

glucose, impaired glucose tolerance, diabetes, and the metabolic syndrome: analysis<br />

of Third National Health and Nutrition Examination Survey (NHANES III).<br />

Ann Epidemiol 14, 686 (2004).<br />

25. Clement, K. et al. Weight loss regulates inflammation-related genes in white<br />

adipose tissue of obese subjects. FASEB J 18, 1657 (2004).<br />

26. Halaas, J.L. et al. Weight-reducing effects of the plasma protein encoded by the<br />

obese gene. Science 269, 543 (1995).<br />

27. Montague, C.T. et al. Congenital leptin deficiency is associated with severe earlyonset<br />

obesity in humans. Nature 387, 903 (1997).<br />

28. Farooqi, I.S. et al. Beneficial effects of leptin on obesity, T cell hypo-responsiveness,<br />

and neuroendocrine/metabolic dysfunction of human congenital leptin deficiency.<br />

J Clin Invest 110, 1093 (2002).<br />

29. Munzberg, H., Flier, J.S., and Bjorbaek, C. Region-specific leptin resistance within<br />

the hypothalamus of diet-induced obese mice. Endocrinology 145, 4880 (2004).<br />

30. Shimomura, I., Hammer, R.E., Ikemoto, S., Brown, M.S., and Goldstein, J.L.<br />

Leptin reverses insulin resistance and diabetes mellitus in mice with congenital<br />

lipodystrophy. Nature 401, 73 (1999).<br />

31. Dunajska, K. et al. Plasma adiponectin concentration in relation to severity of<br />

coronary atherosclerosis and cardiovascular risk factors in middle-aged men.<br />

Endocrine 25, 215 (2004).<br />

32. Yamauchi, T. et al. The fat-derived hormone adiponectin reverses insulin resistance<br />

associated with both lipoatrophy and obesity. Nat Med 7, 941 (2001).<br />

33. Tilg, H. and Wolf, A.M. Adiponectin: a key fat-derived molecule regulating inflammation.<br />

Expert Opin Ther Targets 9, 245 (2005).<br />

34. Koh, K.K et al. Beneficial effects of fenofibrate to improve endothelial dysfunction<br />

and raise adiponectin levels in patients with primary hypertriglyceridemia. Diabetes<br />

Care 28, 1419 (2005).<br />

35. Arita, Y. et al. Paradoxical decrease of an adipose-specific protein, adiponectin,<br />

in obesity. Biochem Biophys Res Commun 257, 79 (1999).<br />

36. Ouchi, N. et al. Novel modulator for endothelial adhesion molecules: adipocytederived<br />

plasma protein adiponectin. Circulation 100, 2473 (1999).<br />

37. Shibata, R. et al. Adiponectin stimulates angiogenesis in response to tissue<br />

ischemia through stimulation of amp-activated protein kinase signaling. J Biol<br />

Chem 279, 28670 (2004).<br />

38. Kern, P.A., Ranganathan, S., Li, C., Wood, L., and Ranganathan, G. Adipose tissue<br />

tumor necrosis factor and interleukin-6 expression in human obesity and insulin<br />

resistance. Am J Physiol Endocrinol Metab 280, E745 (2001).<br />

39. Prins, J.B. et al. Tumor necrosis factor-alpha induces apoptosis of human adipose<br />

cells. Diabetes 46, 1939 (1997).<br />

40. Qian, H. et al. TNFalpha induces and insulin inhibits caspase 3-dependent adipocyte<br />

apoptosis. Biochem Biophys Res Commun 284, 1176 (2001).<br />

41. Zhang, H.H., Kumar, S., Barnett, A.H., and Eggo, M.C. Dexamethasone inhibits<br />

tumor necrosis factor-alpha-induced apoptosis and interleukin-1 beta release in<br />

human subcutaneous adipocytes and preadipocytes. J Clin Endocrinol Metab 86,<br />

2817 (2001).


Obesity and Inflammation: Implications for Atherosclerosis 163<br />

42. Kern, P.A. et al. The expression of tumor necrosis factor in human adipose tissue.<br />

Regulation by obesity, weight loss, and relationship to lipoprotein lipase. J Clin<br />

Invest 95, 2111 (1995).<br />

43. Memon, R.A., Feingold, K.R., Moser, A.H., Fuller, J., and Grunfeld, C. Regulation<br />

of fatty acid transport protein and fatty acid translocase mRNA levels by endotoxin<br />

and cytokines. Am J Physiol 274, E210 (1998).<br />

44. Hotamisligil, G.S., Arner, P., Caro, J.F., Atkinson, R.L., and Spiegelman, B.M.<br />

Increased adipose tissue expression of tumor necrosis factor-alpha in human<br />

obesity and insulin resistance. J Clin Invest 95, 2409 (1995).<br />

45. Koistinen, H.A. et al. Subcutaneous adipose tissue expression of tumour necrosis<br />

factor-alpha is not associated with whole body insulin resistance in obese nondiabetic<br />

or in type-2 diabetic subjects. Eur J Clin Invest 30, 302 (2000).<br />

46. Moller, D.E. Potential role of TNF-alpha in the pathogenesis of insulin resistance<br />

and type 2 diabetes. Trends Endocrinol Metab 11, 212 (2000).<br />

47. Avellone, G. et al. Evaluation of cardiovascular risk factors in overweight and<br />

obese subjects. Int Angiol 13, 25 (1994).<br />

48. Anand, S.S. et al. Relationship of metabolic syndrome and fibrinolytic dysfunction<br />

to cardiovascular disease. Circulation 108, 420 (2003).<br />

49. Mertens, I. and Van Gaal, L.F. Visceral fat as a determinant of fibrinolysis and<br />

hemostasis. Semin Vasc Med 5, 48 (2005).<br />

50. Lyon, C.J. and Hsueh, W.A. Effect of plasminogen activator inhibitor-1 in diabetes<br />

mellitus and cardiovascular disease. Am J Med 115, Suppl 8A, 62S (2003).<br />

51. De Pergola, G. et al. Increase in both pro-thrombotic and anti-thrombotic factors<br />

in obese premenopausal women: relationship with body fat distribution. Int J Obes<br />

Relat Metab Disord 21, 527 (1997).<br />

52. Mertens, I. et al. Visceral fat is a determinant of PAI-1 activity in diabetic and<br />

non-diabetic overweight and obese women. Horm Metab Res 33, 602 (2001).<br />

53. Janand-Delenne, B. et al. Visceral fat as a main determinant of plasminogen<br />

activator inhibitor 1 level in women. Int J Obes Relat Metab Disord 22, 312 (1998).<br />

54. Juhan-Vague, I., Thompson, S.G., and Jespersen, J. Involvement of the hemostatic<br />

system in the insulin resistance syndrome: study of 1500 patients with angina<br />

pectoris. Arterioscler Thromb 13, 1865 (1993).<br />

55. McTernan, P.G., Kusminski, C.M., and Kumar, S. Resistin. Curr Opin Lipidol 17,<br />

170 (2006).<br />

56. Steppan, C.M. et al. A family of tissue-specific resistin-like molecules. Proc Natl<br />

Acad Sci USA 98, 502 (2001).<br />

57. Rajala, M.W., Obici, S., Scherer, P.E., and Rossetti, L. Adipose-derived resistin<br />

and gut-derived resistin-like molecule-beta selectively impair insulin action on<br />

glucose production. J Clin Invest 111, 225 (2003).<br />

58. Patel, L. et al. Resistin is expressed in human macrophages and directly regulated<br />

by PPAR gamma activators. Biochem Biophys Res Commun 300, 472 (2003).<br />

59. Rangwala, S.M. et al. Abnormal glucose homeostasis due to chronic hyperresistinemia.<br />

Diabetes 53, 1937 (2004).<br />

60. Kaser, S. et al. Resistin messenger-RNA expression is increased by pro-inflammatory<br />

cytokines in vitro. Biochem Biophys Res Commun 309, 286 (2003).<br />

61. Shetty, G.K., Economides, P.A., Horton, E.S., Mantzoros, C.S., and Veves, A.<br />

Circulating adiponectin and resistin levels in relation to metabolic factors, inflammatory<br />

markers, and vascular reactivity in diabetic patients and subjects at risk<br />

for diabetes. Diabetes Care 27, 2450 (2004).


164 Oxidative Stress and Inflammatory Mechanisms<br />

62. Ross, R. Atherosclerosis is an inflammatory disease. Am Heart J 138, S419 (1999).<br />

63. Iiyama, K. et al. Patterns of vascular cell adhesion molecule-1 and intercellular<br />

adhesion molecule-1 expression in rabbit and mouse atherosclerotic lesions and<br />

at sites predisposed to lesion formation. Circ Res 85, 199 (1999).<br />

64. van Furth, R. Human monocytes and cytokines. Res Immunol 149, 719 (1998).<br />

65. Fogelman, A.M., Haberland, M.E., Seager, J., Hokom, M., and Edwards, P.A.<br />

Factors regulating the activities of the low density lipoprotein receptor and the<br />

scavenger receptor on human monocyte-macrophages. J Lipid Res 22, 1131<br />

(1981).<br />

66. Heinecke, J.W. Mechanisms of oxidative damage by myeloperoxidase in atherosclerosis<br />

and other inflammatory disorders. J Lab Clin Med 133, 321 (1999).<br />

67. Galis, Z.S., Sukhova, G.K., Lark, M.W., and <strong>Lib</strong>by, P. Increased expression of<br />

matrix metalloproteinases and matrix degrading activity in vulnerable regions of<br />

human atherosclerotic plaques. J Clin Invest 94, 2493 (1994).<br />

68. Osterud, B. Tissue factor expression by monocytes: regulation and pathophysiological<br />

roles. Blood Coagul Fibrinolysis 9, Suppl 1, S9 (1998).<br />

69. Toschi, V. et al. Tissue factor modulates the thrombogenicity of human atherosclerotic<br />

plaques. Circulation 95, 594 (1997).<br />

70. Semplicini, A. et al. Interactions between insulin and sodium homeostasis in<br />

essential hypertension. Am J Med Sci 307, Suppl 1, S43 (1994).<br />

71. Weisberg, S.P. et al. Obesity is associated with macrophage accumulation in<br />

adipose tissue. J Clin Invest 112, 1796 (2003).<br />

72. Stolar, M.W. Insulin resistance, diabetes, and the adipocyte. Am J Health Syst<br />

Pharm 59, Suppl 9, S3 (2002).<br />

73. Xu, H. et al. Chronic inflammation in fat plays a crucial role in the development<br />

of obesity-related insulin resistance. J Clin Invest 112, 1821 (2003).<br />

74. Esposito, K. et al. Effect of weight loss and lifestyle changes on vascular inflammatory<br />

markers in obese women: a randomized trial. JAMA 289, 1799 (2003).<br />

75. Esposito, K. et al. Effect of a mediterranean-style diet on endothelial dysfunction<br />

and markers of vascular inflammation in the metabolic syndrome: a randomized<br />

trial. JAMA 292, 1440 (2004).<br />

76. Liao, J.K. Clinical implications for statin pleiotropy. Curr Opin Lipidol 16, 624<br />

(2005).<br />

77. Ridker, P.M. et al. Inflammation, pravastatin, and the risk of coronary events after<br />

myocardial infarction in patients with average cholesterol levels: Cholesterol and<br />

Recurrent Events (CARE) Investigators. Circulation 98, 839 (1998).<br />

78. Ridker, P.M. et al. Measurement of C-reactive protein for the targeting of statin<br />

therapy in the primary prevention of acute coronary events. New Engl J Med 344,<br />

1959 (2001).<br />

79. Koh, K.K. et al. Comparison of effects of losartan, irbesartan, and candesartan on<br />

flow-mediated brachial artery dilation and on inflammatory and thrombolytic<br />

markers in patients with systemic hypertension. Am J Cardiol 93, 1432, A10<br />

(2004).<br />

80. Wang, C.H. et al. C-reactive protein up-regulates angiotensin type 1 receptors in<br />

vascular smooth muscle. Circulation 107, 1783 (2003).<br />

81. Haffner, S.M. The metabolic syndrome: inflammation, diabetes mellitus, and<br />

cardiovascular disease. Am J Cardiol 97, 3A (2006).


Obesity and Inflammation: Implications for Atherosclerosis 165<br />

82. Haffner, S.M. et al. Effect of rosiglitazone treatment on nontraditional markers of<br />

cardiovascular disease in patients with type 2 diabetes mellitus. Circulation 106,<br />

679 (2002).<br />

83. MRC/BHF Heart Protection Study of cholesterol lowering with simvastatin in<br />

20,536 high-risk individuals: a randomised placebo-controlled trial. Lancet 360,<br />

7 (2002).<br />

84. Colhoun, H.M. et al. Primary prevention of cardiovascular disease with atorvastatin<br />

in type 2 diabetes in the Collaborative Atorvastatin Diabetes Study (CARDS)<br />

multicentre randomised placebo-controlled trial. Lancet 364, 685 (2004).<br />

85. Freeman, D.J. et al. Pravastatin and the development of diabetes mellitus: evidence<br />

for a protective treatment effect in the West of Scotland Coronary Prevention<br />

Study. Circulation 103, 357 (2001).<br />

86. Yusuf, S. et al. Effects of an angiotensin-converting-enzyme inhibitor, ramipril,<br />

on cardiovascular events in high-risk patients: Heart Outcomes Prevention Evaluation<br />

(HOPE) Study Investigators. New Engl J Med 342, 145 (2000).<br />

87. Monzillo, L.U. et al. Effect of lifestyle modification on adipokine levels in obese<br />

subjects with insulin resistance. Obes Res 11, 1048 (2003).<br />

88. Koh, K.K. et al. Comparative effects of diet and statin on NO bioactivity and<br />

matrix metalloproteinases in hypercholesterolemic patients with coronary artery<br />

disease. Arterioscler Thromb Vasc Biol 22, e19 (2002).<br />

89. Orem, C. et al. The effects of atorvastatin treatment on the fibrinolytic system in<br />

dyslipidemic patients. Jpn Heart J 45, 977 (2004).<br />

90. Ludwig, S. et al. Impact of simvastatin on hemostatic and fibrinolytic regulators<br />

in type 2 diabetes mellitus. Diabetes Res Clin Pract 70, 110 (2005).<br />

91. Hoekstra, T., Geleijnse, J.M., Schouten, E.G., and Kluft, C. Plasminogen activator<br />

inhibitor-type 1: its plasma determinants and relation with cardiovascular risk.<br />

Thromb Haemost 91, 861 (2004).


10 Oligomeric<br />

Composition of<br />

Adiponectin and<br />

Obesity<br />

CONTENTS<br />

T. Bobbert and Joachim Spranger<br />

Introduction .......................................................................................................167<br />

Adipocytes and Adiponectin.............................................................................168<br />

Biological Activities of Adiponectin Multimers ..............................................169<br />

Results of Weight Loss .....................................................................................171<br />

Central Nervous System Effects.......................................................................172<br />

Summary ...........................................................................................................173<br />

References .........................................................................................................173<br />

INTRODUCTION<br />

Obesity is among the most frequently encountered metabolic diseases worldwide.<br />

Moreover, its incidence and prevalence are rising rapidly. 1,2 More than half the<br />

world population is considered overweight. 3 Being overweight constitutes a health<br />

risk because it is associated with several co-morbidities including dyslipidemia,<br />

hypertension, type 2 diabetes, and atherosclerotic cardiovascular disease. 4.5 Adipose<br />

tissue was initially believed to be only a fat storage organ, but it is now<br />

acknowledged to be an active participant in energy homeostasis and other physiological<br />

functions. Adipose tissue is known to express and secrete a variety of<br />

novel adipocytokines that have been implicated in the development of insulin<br />

resistance and atherosclerosis. 6,7 Dysregulation of adipocytokine production is<br />

directly involved in the pathophysiology of metabolic syndrome, and normalization<br />

of plasma concentrations of adipocytokines reverses the phenotype of metabolic<br />

syndrome. 8,9<br />

167


168 Oxidative Stress and Inflammatory Mechanisms<br />

ADIPOCYTES <strong>AND</strong> ADIPONECTIN<br />

Adipocytes secrete a variety of polypeptides such as leptin, resistin, and adiponectin.<br />

Adiponectin, the gene product of the adipose tissue’s most abundant<br />

gene transcript, 10 may be a link between obesity and the development of insulin<br />

resistance. cDNAs for adiponectin were first identified independently in different<br />

mouse cells 11,12 and by large-scale random sequencing of a 3′-directed human<br />

adipose tissue cDNA library. 10 Subsequently, adiponectin was purified from<br />

human plasma, 13 where it circulates in considerable concentrations. Especially in<br />

the regulation of the glucose and lipid metabolism, adiponectin has been shown<br />

to play an important role. 11–13<br />

Unlike other adipose-derived proteins, plasma levels of adiponectin have been<br />

found to be decreased in a number of deranged metabolic states including obesity,<br />

14 dyslipidemia, 15 type 2 diabetes, and insulin resistance. 12,16,17 Many studies<br />

have demonstrated that reduced circulating adiponectin levels can be partially<br />

elevated after induction of weight loss in obese and insulin-resistant subjects. 17,18<br />

Apart from a connection to obesity or diabetes, further parameters like age, sex<br />

hormones, and glucocorticoids may play parts in the regulation of adiponectin<br />

levels. 19,20<br />

Adiponectin is composed of a carboxyl-terminal globular domain and an<br />

amino-terminal collagenous domain. 21,22 It belongs to the soluble collagen superfamily<br />

and has structural homology with collagen VIII, X, complement factor<br />

C1q, 16 and the tumor necrosis (TNF) family. 10,21 This kind of structure is known<br />

to form characteristic multimers. 23,24<br />

Gel filtration and velocity gradient sedimentation studies revealed adiponectin<br />

circulating in serum to form several different molecular weight species; the largest<br />

species was more than several hundred kilodaltons in size. 11,13,14 Scherer et al.<br />

noted that adiponectin from 3T3-L1 adipocytes forms trimers, hexamers, and<br />

larger multimers. 11 Tsao et al. and Arita et al. analyzed multimer formation of<br />

adiponectin in serum by gel filtration chromatography and showed adiponectin<br />

to be separable into three species. 14,25<br />

Kadowaki et al. showed a method of evaluating the multimer formation of<br />

adiponectin. 26 Using SDS-PAGE under non-reducing and non-heat-denaturing<br />

conditions, they separated multimers of adiponectin from various sources into<br />

three species; LMW trimers, MMW multimers, and HMW multimers. This classification<br />

is not used in all publications and depends partially on the determination<br />

method used. Also the MMW fractions or hexamers are added to the LMW<br />

oligomers and predictions or correlations were calculated by the ratio of HMW<br />

to MMW plus LMW species.<br />

Oligomer formation of adiponectin depends on the disulfide bond formation<br />

mediated by Cys-39. 27 Adiponectin was reported to be an α-2,8-linked disialic<br />

acid-containing glycoprotein, although the biological functions of the disialic acid<br />

epitope of adiponectin remain to be elucidated. 28 The regulation of adiponectin<br />

multimerization and secretion occurs also via changes in post-translational modifications<br />

(PTMs). PTMs identified in murine and bovine adiponectin include


Oligomeric Composition of Adiponectin and Obesity 169<br />

hydroxylation of multiple conserved proline and lysine residues and glycosylation<br />

of hydroxylysines.<br />

BIOLOGICAL ACTIVITIES OF ADIPONECTIN MULTIMERS<br />

Discussion of the biological activities of the different adiponectin multimers has<br />

proven controversial. HMW multimer levels appear to be higher in women compared<br />

to men 26 and adiponectin exerts multiple metabolic actions at multiple tissue<br />

sites. The isolated globular domain of adiponectin stimulates fatty acid oxidation<br />

in skeletal muscle, whereas full length adiponectin synergizes with insulin to<br />

inhibit hepatic glucose production. 16,29,30 In mice, disruption of the adiponectin<br />

locus leading to its ablation resulted in impaired fatty acid clearance, increased<br />

tumor necrosis factor-α levels, and aggravated insulin resistance in animals fed<br />

high fat diets. 31,32<br />

The HMW and hexameric adiponectin can activate transcription factor NFκB<br />

in undifferentiated or differentiated C2C12 cells, but trimeric adiponectin and<br />

the isolated globular domain of adiponectin cannot. The isolated globular domain<br />

of adiponectin, but not full-length adiponectin hexamer, enhances muscle fatty<br />

acid oxidation by inactivating acetyl-CoA carboxylase following stimulation of<br />

AMP-activated protein kinase (AMPK). 33,34 Waki et al. reported specifically that<br />

the HMW isoform promotes AMP-activated protein kinase in hepatocytes. 26 In<br />

contrast, Tsao et al. recently reported that only trimers activate AMPK in muscle,<br />

whereas hexamers and HMW form activated NF-κB. 35<br />

Differences in the tissue-specific expression patterns of two adiponectin<br />

receptors may contribute to these divergent activities. 36 Bobbert et al. investigated<br />

the effects of moderate weight reduction by a lifestyle intervention on adiponectin<br />

oligomer composition and its relation to glucose and fat metabolism. 37 While<br />

HMW and MMW adiponectins increased, a decreased amount of LMW adiponectin<br />

was found after weight reduction. Total adiponectin and especially HMW<br />

adiponectin correlated strongly with HDL cholesterol while correlations with<br />

other markers of glucose or fat metabolism were weak.<br />

Many studies have demonstrated a correlation between total adiponectin and<br />

HDL cholesterol. Most studies suggested that hypoadiponectinemia is more<br />

closely related to adiposity and dyslipidemia rather than insulin sensitivity. 38<br />

Indeed, Bobbert et al. found a strong correlation between HMW adiponectin and<br />

HDL cholesterol, which suggests that the relationship between total adiponectin<br />

and HDL cholesterol is primarily driven by HMW adiponectin rather than total<br />

adiponectin (Figure 10.1).<br />

HDL cholesterol is basically generated from lipid-free apolipoprotein A-I or<br />

lipid-poor pre-ß1-HDL as a precursor. These precursors are partially produced<br />

by the liver and it is well known that adiponectin oligomers specifically affect<br />

liver metabolism. 25 The importance of considering adiponectin oligomers is supported<br />

by a multivariate analysis revealing that HMW adiponectin explained about<br />

30% of the variability of HDL cholesterol (Figure 10.2). These results confirm<br />

those of Baratta and co-workers, who demonstrated that adiponectin is correlated


170 Oxidative Stress and Inflammatory Mechanisms<br />

a)<br />

b)<br />

HDL (mmol/l)<br />

HDL (mmol/l)<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

Adiponectin (μg/ml)<br />

r = 0.61 (p = 0.016)<br />

0 2 4 6 8 10 12<br />

r = 0.665 (p = 0.007)<br />

0.0<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4<br />

HMW absolute (μg/ml)<br />

FIGURE 10.1 Correlation between (a) total adiponectin and HDL and (b) absolute concentrations<br />

of adiponectin oligomers and HDL in participants of a weight reduction<br />

program after moderate weight loss.


Oligomeric Composition of Adiponectin and Obesity 171<br />

with serum lipid improvement independently of insulin sensitivity changes after<br />

weight loss. 39<br />

p38MAP kinase<br />

Glc uptake<br />

Skeletal muscle<br />

AdipoR1 AdipoR2<br />

P CBS<br />

α γ<br />

AMPK β<br />

FA oxidation<br />

Skeletal muscle<br />

FIGURE 10.2 Adiponectin-dependent intracellular pathways.<br />

Different adiponectin oligomers and the varying appearances of the AdipoR1<br />

and AdipoR2 adiponectin receptors may be responsible for the different actions<br />

of adiponectin oligomers on fat or glucose metabolism. However, the precise<br />

molecular mechanism for the oligomer-specific effect is still unclear. The mechanism<br />

that regulates adiponectin oligomer composition has not been identified.<br />

Data from intervention studies with thiazolidinediones suggest that the process<br />

is PPAR-γ-dependent. 40 Because PPAR-γ is activated by negative energy balance<br />

and weight reduction, 41 the effect on adiponectin oligomers may also depend on<br />

PPAR-γ-related pathways. Tsuchida et al. showed the influence of PPAR-α and<br />

γ on the expression of adiponectin and AdipoR1 and AdipoR2. They showed that<br />

activation of PPAR-γ or food restriction increased the ratio of HMW to total<br />

adiponectin and that activation of PPARα did not affect the ratio.<br />

RESULTS OF WEIGHT LOSS<br />

PPARα<br />

FA oxidation<br />

Liver<br />

The results of Bobbert et al. are in agreement with other studies showing that<br />

moderate weight loss results in relatively small changes of circulating adiponectin<br />

levels. 42 One study of six patients and a 3-month follow-up investigated the effects


172 Oxidative Stress and Inflammatory Mechanisms<br />

of weight reduction on adiponectin oligomers. This study primarily aimed to<br />

investigate the effects of HMW adiponectin on endothelial cell apoptosis. Due<br />

to the small number of participants, the precise relationship of adiponectin oligomers<br />

to metabolic changes after weight loss was not further evaluated. 43<br />

Physical activity correlates strongly with obesity and is apart from dietary<br />

interventions a major part of weight reduction. Physical exercise is associated<br />

with reduced risks for the development of obesity-associated co-morbidities like<br />

type 2 diabetes mellitus 44 and also reduces the mortality risks of individuals with<br />

impaired glucose tolerance to the levels of healthy persons. 45 The improvement<br />

of insulin sensitivity by physical activity has been proposed as a possible mechanism<br />

of these effects. 46 From a mechanistic view, adipokines have been identified<br />

as potential mediators between obesity and insulin sensitivity. Despite the temptation<br />

to speculate that circulating adiponectin may be affected by degrees of<br />

physical activity, most studies have found that physical exercise has no influence<br />

on total adiponectin levels. Neither short term exercise nor long term physical<br />

training exerted effects on total adiponectin plasma levels. 47,48 Rather controversially,<br />

a study by Jurimae et al. reported reductions directly after acute rowing.<br />

However, 30 min after the rowing levels increased above resting values, 49,50<br />

Kriketos et al. observed increased adiponectin levels following two or three bouts<br />

of exercise over 1 week. These values remained elevated after 10 weeks of<br />

exercise. 49,50 Bobbert et al. showed that total adiponectin and oligomers were<br />

unchanged by acute and chronic exercise (in press). Total adiponectin levels and<br />

oligomers were not different for trained or untrained persons and were also<br />

unaltered by different training intensities. It is therefore unlikely that changes of<br />

adiponectin oligomers are responsible for the beneficial effects of sustained<br />

physical exercise on lipid and glucose metabolism. However, total adiponectin<br />

and again specifically HMW oligomers correlated with HDL cholesterol in this<br />

study, which supported the hypothesis that HMW adiponectin specifically mediates<br />

the positive effects of adiponectin.<br />

CENTRAL NERVOUS SYSTEM EFFECTS<br />

In addition to the peripheral influence of adiponectin on human metabolism,<br />

central nervous effects also play an important role in the regulation of human<br />

energy balance and metabolism. A growing body of evidence suggests that adiponectin<br />

directly affects energy balance by increasing thermogenesis. 51 Adiponectin<br />

has recently been reported to generate a negative energy balance by increasing<br />

energy expenditure. 52<br />

Spranger et al. showed that neither radiolabeled non-glycosylated nor glycosylated<br />

globular adiponectin crossed the blood–brain barrier (BBB) in mice. 53<br />

In addition, adiponectin and adiponectin oligomers were not detectable in human<br />

cerebrospinal fluid using various established methods. Using murine cerebral<br />

microvessels, they demonstrated expression of adiponectin receptors that were<br />

up-regulated during fasting in brain endothelium. Interestingly, treatment with<br />

adiponectin reduced secretion of centrally active interleukin-6 from brain


Oligomeric Composition of Adiponectin and Obesity 173<br />

endothelial cells — a phenomenon paralleled by similar trends of other<br />

pro-inflammatory cytokines. These data suggest that direct effects of endogenous<br />

adiponectin on central nervous system pathways are unlikely to exist. However,<br />

the identification of adiponectin receptors on brain endothelial cells and the<br />

finding of a modified secretion pattern of centrally active substances from BBB<br />

cells provide an alternate explanation as to how adiponectin may evoke effects<br />

on energy metabolism.<br />

SUMMARY<br />

Adiponectin plays an important role in human metabolism, although the exact<br />

function of adiponectin is still unclear. The influences of adiponectin and adiponectin<br />

oligomers on peripheral glucose and lipid metabolism serve as foci of<br />

current clinical research. Some in vitro analyses showed first hints of adiponectin<br />

oligomer-dependent intracellular signalling cascades and initial human data demonstrated<br />

the different peripheral and central nervous system actions of total<br />

adiponectin and the different adiponectin oligomers.<br />

REFERENCES<br />

1. Caterson ID and Gill TP. Obesity: epidemiology and possible prevention. Best<br />

Pract Res Clin Endocrinol Metab 2002; 16: 595.<br />

2. Formiguera X and Canton A. Obesity: epidemiology and clinical aspects. Best<br />

Pract Res Clin Gastroenterol 2004; 18: 1125.<br />

3. Field AE et al. Impact of overweight on the risk of developing common chronic<br />

diseases during a 10-year period. Arch Intern Med 2001; 161: 1581.<br />

4. Linton MF and Fazio S. A practical approach to risk assessment to prevent<br />

coronary artery disease and its complications. Am J Cardiol 2003; 92: 19i.<br />

5. Scott CL. Diagnosis, prevention, and intervention for the metabolic syndrome.<br />

Am J Cardiol 2003; 92: 35i.<br />

6. Matsuzawa Y, Funahashi T, and Nakamura T. Molecular mechanism of metabolic<br />

syndrome X: contribution of adipocytokines adipocyte-derived bioactive substances.<br />

Ann NY Acad Sci 1999; 892: 146.<br />

7. Funahashi T et al. Role of adipocytokines on the pathogenesis of atherosclerosis<br />

in visceral obesity. Intern Med 1999; 38: 202.<br />

8. Hotamisligil GS, Shargill NS, and Spiegelman BM. Adipose expression of tumor<br />

necrosis factor-alpha: direct role in obesity-linked insulin resistance. Science 1993;<br />

259: 87.<br />

9. Shimomura I et al. Leptin reverses insulin resistance and diabetes mellitus in mice<br />

with congenital lipodystrophy. Nature 1999; 401: 73.<br />

10. Maeda K, Okubo K, Shimomura I, Funahashi T, Matsuzawa Y, and Matsubara K.<br />

cDNA cloning and expression of a novel adipose specific collagen-like factor,<br />

apM1 (adipose most abundant gene transcript 1). Biochem Biophys Res Commun<br />

1996; 221: 286.


174 Oxidative Stress and Inflammatory Mechanisms<br />

11. Scherer PE, Williams S, Fogliano M, Baldini G, and Lodish HF. A novel serum<br />

protein similar to C1q produced exclusively in adipocytes. J Biol Chem 1995;<br />

270: 26746.<br />

12. Hu E, Liang P, and Spiegelman BM. AdipoQ is a novel adipose-specific gene<br />

dysregulated in obesity. J Biol Chem 1996; 271: 10697.<br />

13. Nakano Y, Tobe T, Choi-Miura NH, Mazda T, and Tomita M. Isolation and<br />

characterization of GBP28, a novel gelatin-binding protein purified from human<br />

plasma. J Biochem (Tokyo) 1996; 120: 803.<br />

14. Arita Y et al. Paradoxical decrease of an adipose-specific protein, adiponectin, in<br />

obesity. Biochem Biophys Res Commun 1999; 257: 79.<br />

15. Matsubara M, Maruoka S, and Katayose S. Decreased plasma adiponectin concentrations<br />

in women with dyslipidemia. J Clin Endocrinol Metab 2002; 87: 2764.<br />

16. Yamauchi T et al. The fat-derived hormone adiponectin reverses insulin resistance<br />

associated with both lipoatrophy and obesity. Nat Med 2001; 7: 941.<br />

17. Hotta K et al. Plasma concentrations of a novel, adipose-specific protein, adiponectin,<br />

in type 2 diabetic patients. Arterioscler Thromb Vasc Biol 2000; 20: 1595.<br />

18. Yang WS et al. Weight reduction increases plasma levels of an adipose-derived<br />

anti-inflammatory protein, adiponectin. J Clin Endocrinol Metab 2001; 86: 3815.<br />

19. Reinehr T, Roth C, Menke T, and Andler W. Adiponectin before and after weight<br />

loss in obese children. J Clin Endocrinol Metab 2004; 89: 3790.<br />

20. Diez JJ and Iglesias P. The role of the novel adipocyte-derived hormone adiponectin<br />

in human disease. Eur J Endocrinol 2003; 148: 293.<br />

21. Shapiro L and Scherer PE. The crystal structure of a complement-1q family protein<br />

suggests an evolutionary link to tumor necrosis factor. Curr Biol 1998; 8: 335.<br />

22. Yokota T et al. Adiponectin, a new member of the family of soluble defense<br />

collagens, negatively regulates the growth of myelomonocytic progenitors and the<br />

functions of macrophages. Blood 2000; 96: 1723.<br />

23. Crouch E, Persson A, Chang D, and Heuser J. Molecular structure of pulmonary<br />

surfactant protein D (SP-D). J Biol Chem 1994; 269: 17311.<br />

24. McCormack FX et al. The Cys6 intermolecular disulfide bond and the collagenlike<br />

region of rat SP-A play critical roles in interactions with alveolar type II cells<br />

and surfactant lipids. J Biol Chem 1997; 272: 27971.<br />

25. Tsao TS, Murrey HE, Hug C, Lee DH, and Lodish HF. Oligomerization statedependent<br />

activation of NF-kappa B signaling pathway by adipocyte complementrelated<br />

protein of 30 kDa (Acrp30). J Biol Chem 2002; 277: 29359.<br />

26. Waki H et al. Impaired multimerization of human adiponectin mutants associated<br />

with diabetes. Molecular structure and multimer formation of adiponectin. J Biol<br />

Chem 2003; 278: 40352.<br />

27. Pajvani UB et al. Structure-function studies of the adipocyte-secreted hormone<br />

Acrp30/adiponectin: implications for metabolic regulation and bioactivity. J Biol<br />

Chem 2003; 278: 9073.<br />

28. Sato C, Yasukawa Z, Honda N, Matsuda T, and Kitajima K. Identification and<br />

adipocyte differentiation-dependent expression of the unique disialic acid residue<br />

in an adipose tissue-specific glycoprotein, adipo Q. J Biol Chem 2001; 276: 28849.<br />

29. Berg AH, Combs TP, Du X, Brownlee M, and Scherer PE. The adipocyte-secreted<br />

protein Acrp30 enhances hepatic insulin action. Nat Med 2001; 7: 947.<br />

30. Fruebis J et al. Proteolytic cleavage product of 30-kDa adipocyte complementrelated<br />

protein increases fatty acid oxidation in muscle and causes weight loss in<br />

mice. Proc Natl Acad Sci USA 2001; 98: 2005.


Oligomeric Composition of Adiponectin and Obesity 175<br />

31. Maeda N et al. Diet-induced insulin resistance in mice lacking adiponectin/ACRP30.<br />

Nat Med 2002; 8: 731.<br />

32. Kubota N et al. Disruption of adiponectin causes insulin resistance and neointimal<br />

formation. J Biol Chem 2002; 277: 25863.<br />

33. Tomas E et al. Enhanced muscle fat oxidation and glucose transport by ACRP30<br />

globular domain: acetyl-CoA carboxylase inhibition and AMP-activated protein<br />

kinase activation. Proc Natl Acad Sci USA 2002; 99: 16309.<br />

34. Yamauchi T et al. Adiponectin stimulates glucose utilization and fatty-acid oxidation<br />

by activating AMP-activated protein kinase. Nat Med 2002; 8: 1288.<br />

35. Tsao TS et al. Role of disulfide bonds in Acrp30/adiponectin structure and signaling<br />

specificity: different oligomers activate different signal transduction pathways.<br />

J Biol Chem 2003; 278: 50810.<br />

36. Yamauchi T et al. Cloning of adiponectin receptors that mediate antidiabetic<br />

metabolic effects. Nature 2003; 423: 762.<br />

37. Bobbert T et al. Changes of adiponectin oligomer composition by moderate weight<br />

reduction. Diabetes 2005; 54: 2712.<br />

38. Kazumi T, Kawaguchi A, Hirano T, and Yoshino G. Serum adiponectin is associated<br />

with high-density lipoprotein cholesterol, triglycerides, and low-density lipoprotein<br />

particle size in young healthy men. Metabolism 2004; 53: 589.<br />

39. Baratta R et al. Adiponectin relationship with lipid metabolism is independent of<br />

body fat mass: evidence from both cross-sectional and intervention studies. J Clin<br />

Endocrinol Metab 2004; 89: 2665.<br />

40. Pajvani UB et al. Complex distribution, not absolute amount of adiponectin,<br />

correlates with thiazolidinedione-mediated improvement in insulin sensitivity. J<br />

Biol Chem 2004; 279: 12152.<br />

41. Verreth W et al. Weight loss-associated induction of peroxisome proliferatoractivated<br />

receptor-alpha and peroxisome proliferator-activated receptor-gamma<br />

correlate with reduced atherosclerosis and improved cardiovascular function in<br />

obese insulin-resistant mice. Circulation 2004; 110: 3259.<br />

42. Abbasi F et al. Plasma adiponectin concentrations do not increase in association<br />

with moderate weight loss in insulin-resistant, obese women. Metabolism 2004;<br />

53: 280.<br />

43. Kobayashi H et al. Selective suppression of endothelial cell apoptosis by the high<br />

molecular weight form of adiponectin. Circ Res 2004; 94: e27.<br />

44. Helmrich SP, Ragland DR, Leung RW, and Paffenbarger RS, Jr. Physical activity<br />

and reduced occurrence of non-insulin-dependent diabetes mellitus. New Engl J<br />

Med 1991; 325: 147.<br />

45. Eriksson KF and Lindgarde F. No excess 12-year mortality in men with impaired<br />

glucose tolerance who participated in the Malmo Preventive Trial with diet and<br />

exercise. Diabetologia 1998; 41: 1010.<br />

46. Kirwan JP and del Aguila LF. Insulin signalling, exercise and cellular integrity.<br />

Biochem Soc Trans 2003; 31, Pt 6: 1281.<br />

47. Hulver MW et al. Adiponectin is not altered with exercise training despite<br />

enhanced insulin action. Am J Physiol Endocrinol Metab 2002; 283: E861.<br />

48. Kraemer RR et al. Adiponectin responses to continuous and progressively intense<br />

intermittent exercise. Med Sci Sports Exerc 2003; 35: 1320.<br />

49. Kriketos AD et al. Exercise increases adiponectin levels and insulin sensitivity in<br />

humans. Diabetes Care 2004; 27: 629.


176 Oxidative Stress and Inflammatory Mechanisms<br />

50. Jurimae J, Purge P, and Jurimae T. Adiponectin is altered after maximal exercise<br />

in highly trained male rowers. Eur J Appl Physiol 2005; 93: 502.<br />

51. Seeley RJ, D’Alessio DA, and Woods SC. Fat hormones pull their weight in the<br />

CNS. Nat Med 2004; 10: 454.<br />

52. Qi Y et al. Adiponectin acts in the brain to decrease body weight. Nat Med 2004;<br />

10: 524.<br />

53. Spranger J et al. Adiponectin does not cross the blood-brain barrier but modifies<br />

cytokine expression of brain endothelial cells. Diabetes 2006; 55: 141.


11 Insulin-Stimulated<br />

Reactive Oxygen<br />

Species and Insulin<br />

Signal Transduction<br />

CONTENTS<br />

Barry J. Goldstein, Kalyankar Mahadev, and<br />

Xiangdong Wu<br />

Introduction .......................................................................................................178<br />

Regulation of Reversible Tyrosine Phosphorylation in Insulin<br />

Signaling by Tyrosine Phosphatases .......................................................178<br />

ROS as Second Messengers for Cellular Tyrosine Kinase Signaling .............179<br />

Novel Regulatory Paradigm: PTPs Are Thiol-Dependent Enzymes<br />

Regulated by Cellular ROS.....................................................................180<br />

Generation of H 2O 2 by Cellular Insulin Stimulation .......................................181<br />

Insulin-Stimulated H 2O 2 Generation Negatively Regulates PTP1B ................182<br />

Dynamics of Receptor-Induced ROS, PTP Inhibition, and<br />

Signal Transduction.................................................................................183<br />

Identification of the NADPH Oxidase Homolog Nox4 as a Potential<br />

Mediator of Insulin-Stimulated ROS ......................................................184<br />

Regulation of Nox4 Signaling in Insulin Action Cascade...............................184<br />

Potential Role of Rac in Insulin-Stimulated H 2O 2 and PTP Regulation .........185<br />

Gαi2.........................................................................................................185<br />

Novel Targets of Insulin-Stimulated ROS That May Potentially<br />

Influence Insulin Action ..........................................................................186<br />

Summary ...........................................................................................................186<br />

Acknowledgment...............................................................................................187<br />

References .........................................................................................................187<br />

177


178 Oxidative Stress and Inflammatory Mechanisms<br />

INTRODUCTION<br />

Cellular reactive oxygen species (ROS; superoxide and H 2O 2), especially when<br />

chronically raised to high levels and associated with hyperglycemia, are widely<br />

recognized to play an important pathophysiological role in the chronic complications<br />

of diabetes as well as in the development of the disease. 1–3 In contrast,<br />

the transient generation of smaller amounts of ROS is triggered in cells in response<br />

to stimulation with a variety of growth factors, cytokines, and hormones including<br />

insulin, and facilitates their respective signaling cascades. The involvement of an<br />

oxidation step in the action of insulin has been suggested for decades, but only<br />

recently have potential molecular mechanisms been identified for these effects.<br />

Among the signaling enzymes potentially susceptible to inhibition by biochemical<br />

oxidation are those that contain reduced cysteine thiol side chains<br />

essential for their catalytic activities, including the family of protein–tyrosine<br />

phosphatases (PTPs) and other important signal regulators. Recently, we identified<br />

a role for the NADPH oxidase homolog known as Nox4 in the rapid generation<br />

of ROS in insulin-stimulated cells. 4–6 A full understanding of this signaling<br />

network may potentially provide a novel means of facilitating insulin action in<br />

states of insulin resistance and differentially regulating some of the pleiotropic<br />

cellular actions of insulin.<br />

REGULATION OF REVERSIBLE TYROSINE PHOSPHORYLATION<br />

IN INSULIN SIGNALING BY TYROSINE PHOSPHATASES<br />

Insulin is a critical regulator of pleiotropic cellular responses and resistance to<br />

the action of insulin in peripheral tissues is a fundamental defect in type 2<br />

diabetes. 7,8 Insulin action is initiated by binding to a specific plasma membrane<br />

receptor that encodes a tyrosine-specific protein kinase that autophosphorylates<br />

the receptor and its substrate proteins in cells. 9,10 These phosphorylated tyrosine<br />

motifs act as docking scaffolds for the binding and activation of a variety of<br />

signaling and adaptor proteins that are linked to downstream insulin responses.<br />

Specific PTPs regulate the steady-state balance of reversible protein–tyrosine<br />

phosphorylation in the insulin signaling cascade, in concert with the insulin<br />

receptor kinase. 5 In addition to serving as steady-state regulators, PTPs appear<br />

to be required for receptor deactivation since purified insulin receptors retain their<br />

autophosphorylation state after insulin is removed from the ligand binding site<br />

in vitro, and in vivo, dissociation of insulin from the receptor is followed by its<br />

rapid dephosphorylation and deactivation of its kinase activity. 11<br />

In particular, PTP1B has become an important target for therapeutic intervention<br />

in disease states associated with clinical insulin resistance. 12–14 This<br />

single-domain intracellular PTP has been convincingly shown in two knock-out<br />

mouse models to negatively regulate the insulin action cascade in vivo. 15,16 The<br />

cellular basis of these in vivo findings is well documented. 12 PTP1B is active in<br />

vitro against the autophosphorylated insulin receptor 17,18 and it also has a relatively<br />

high specific activity toward IRS-1 compared to other candidate PTPs. 19,20 Several


Oxygen Species and Signal Transduction 179<br />

studies have characterized the unique molecular interactions underlying the close<br />

interaction between the insulin receptor and PTP1B that is facilitated by the<br />

presence of a second phosphotyrosine binding site in the PTP1B catalytic region<br />

that interacts with the multiple phosphotyrosine residues of the receptor<br />

kinase. 21–23<br />

The PTPs comprise an extensive group of homologous proteins involved in<br />

a variety of signal transduction pathways. 24 Receptor and non-receptor forms of<br />

the classical tyrosine-specific PTPs have in common a ~230 amino acid phosphatase<br />

domain that contains the tightly conserved signature catalytic motif<br />

VHCSxGxGR[T/S]G. 25 The non-classical, dual-specificity (tyrosine and serine/<br />

threonine) phosphatases (e.g., MAP kinase phosphatase and others) 26 and PTEN<br />

(phosphatase and tensin homolog deleted on chromosome 10), which dephosphorylates<br />

the 3-phosphate of inositol phospholipids generated by PI 3-kinase<br />

and exhibits dual phosphatase activity in vitro are structurally related, sharing a<br />

less tightly conserved catalytic motif that retains the essential C(x) 5R core structure.<br />

27 This sequence contains the reduced cysteine residue required for catalysis<br />

that is involved in the formation of a cysteinyl-phosphate intermediate. 28,29 Modification<br />

of this catalytic cysteine by oxidation or disulfide conjugation is a<br />

critically important mode of reversible and irreversible PTP regulation in vivo,<br />

including in the insulin action cascade. 30–32<br />

ROS AS SECOND MESSENGERS FOR CELLULAR TYROSINE<br />

KINASE SIGNALING<br />

Superoxide and H 2O 2 are now well recognized to play an integral role in several<br />

growth factor and cytokine signal transduction pathways (recently reviewed in<br />

Rhee). 33 Superoxide [O 2• – ], hydroxyl [•OH] ions, and H 2O 2 generated by cellular<br />

redox reactions have complex physiologies and are ultimately converted to H 2O<br />

+ O 2 by cellular catalase, thioredoxin, glutathione peroxidase, and/or peroxiredoxins.<br />

Relatively low levels of H 2O 2 generated in response to growth factor stimulation<br />

occur in a concerted fashion with specific signaling targets in cells, suggesting<br />

a role as a second messenger. H 2O 2 activates tyrosine phosphorylation<br />

cascades in cultured cells in a manner that mimics ligand-mediated signaling by<br />

PDGF and EGF. However, intracellular H 2O 2 generated transiently during stimulation<br />

of cells with growth factors has been convincingly demonstrated in seminal<br />

experiments by the groups of Finkel et al. 34 and Rhee et al., 35 showing that<br />

autophosphorylation of PDGF and EGF receptors, respectively, and their distal<br />

signaling effects are dependent on post-receptor intracellular H 2O 2 production,<br />

not simply the addition of exogenous H 2O 2. As described below, numerous studies<br />

support the hypothesis that cellular oxidant signaling is mediated by discrete,<br />

localized redox circuitry, distinct from the notion of a generalized “oxidative<br />

stress” effect. 36


180 Oxidative Stress and Inflammatory Mechanisms<br />

NOVEL REGULATORY PARADIGM: PTPS ARE THIOL-DEPENDENT<br />

ENZYMES REGULATED BY CELLULAR ROS<br />

In parallel with developments in the cellular physiology of H 2O 2 generation,<br />

studies have also characterized the biochemical inhibition of PTPs by progressive<br />

oxidation of their catalytic cysteine thiol moieties by cellular ROS to more inert<br />

forms in vivo (Figure 11.1). 30,32,37–39 The activity of PTP1B is dependent on the<br />

oxidation state of its cys-215 residue, which is required for catalytic activity via<br />

the formation of a phospho-enzyme intermediate. 29,37,40,41<br />

The catalytic cysteine residue in the PTP active site is particularly sensitive<br />

to oxidation because of hydrogen bonding of neighboring side chains, which<br />

lowers the thiol p Ka to ~5.5, more than 3 units below that of a typical –SH group,<br />

rendering it in an ionized state at physiological pH. Compared to other typical<br />

protein sulfhydryl side chains, the catalytic PTP thiol can be readily oxidized by<br />

PTP-SSG<br />

(glutathiolated)<br />

inactive<br />

GSH<br />

GSH ?<br />

GSH<br />

GSSG<br />

PTP-S- PTP-S- active<br />

O•<br />

PTP-SOH<br />

(sulfenic acid) O•<br />

inactive<br />

PTP-SN<br />

(sulfenyl-amide)<br />

inactive<br />

PTP-S- PTP-S- reactivated<br />

thiol<br />

reduction<br />

intramolecular<br />

rearrangement<br />

thiol<br />

reduction<br />

PTP-SO 2 H PTP-SO 3 H<br />

inactive, irreversible<br />

FIGURE 11.1 Regulation of PTP catalytic activity by oxidation, reduction, and conjugation.<br />

The catalytic cysteine residue of PTPs is especially reactive because of the low p Ka<br />

of the sulfhydryl that favors a relatively ionized state of the cysteinyl hydrogen. 42 When<br />

subjected to ROS, including those elicited by cellular insulin stimulation, the cysteine side<br />

chain undergoes step-wise oxidation to increasingly inert forms. 32,38,106 The inactive<br />

sulfenic derivative may be reduced to regenerate the active thiol form of the protein.<br />

Alternatively, it may be directly conjugated with glutathione in the cell, producing a<br />

catalytically inert PTP derivative that can be reactivated by biochemical reduction or<br />

through the action of glutathione reductases. 46 Recently, the sulfenic derivative of PTP1B<br />

has been shown to undergo an intramolecular rearrangement, forming a novel sulfenyl–amide<br />

derivative that also sequesters the PTP in an inactive state. 47,48 The sulfenyl–amide<br />

form may actually be an obligate intermediate in this reaction scheme because<br />

its altered protein conformation opens a groove adjacent to the catalytic center that may<br />

render it particularly susceptible to reduction with cytosolic glutathione compared with<br />

the sulfenic derivative. (Reprinted from Goldstein, B.J. et al. Diabetes 54, 311, 2005. With<br />

permission from the American Diabetes Association.)


Oxygen Species and Signal Transduction 181<br />

locally generated H 2O 2 even in the presence of high cytosolic concentrations<br />

(millimolar) of the cysteine-containing tripeptide glutathione (GSH). 42<br />

The catalytic cysteine thiol is initially oxidized to the sulfenic (–SOH) form<br />

that can be reversed by cellular enzymatic mechanisms or with reducing agents<br />

in vitro (Figure 11.1). 40,43 Sequential steps of progressive oxidation to the sulfinic<br />

(–SO 2H) and sulfonic (–SO 3H) forms can lead to irreversible PTP inactivation.<br />

41,44,45 However, the partially oxidized sulfenic acid intermediate of PTP1B<br />

can also be rapidly converted to other forms that may stabilize the molecule and<br />

protect it from further irreversible oxidation.<br />

One potentially stabilizing modification is conjugation with glutathione which<br />

may be enzymatically reactivated in cells by glutaredoxin. 46 The catalytic cysteine<br />

of PTP1B has recently been shown to be reversibly converted to a previously<br />

unknown intramolecular sulfenyl–amide species, in which it becomes linked to<br />

the main chain nitrogen of an adjacent residue, rendering the enzyme inactive<br />

and inducing large conformational changes that inhibit substrate binding. 47,48 This<br />

novel protein modification not only protects the enzyme catalytic site from irreversible<br />

oxidation to sulfonic acid but also permits redox regulation of the enzyme<br />

by promoting its reversible reduction by thiols. The conformation of the catalytic<br />

cleft assumes a more open structure in the sulfenyl–amide derivative of PTP1B<br />

and renders it particularly amenable to reduction by GSH. 47,48 This suggests that<br />

this unique protein derivative may, in fact, be an obligatory intermediate in the<br />

generation of the glutathionylated form of the oxidized enzyme.<br />

GENERATION OF H 2O 2 BY CELLULAR INSULIN STIMULATION<br />

The potential involvement of oxidant species in insulin signaling was initially<br />

explored more than 30 years ago, with the observation by Czech et al. that certain<br />

metal cations interacting with albumin could transfer electrons to a cellular target<br />

and enhance glucose utilization by adipocytes. 49–51 Livingston et al. also showed<br />

that polyamines and related insulin mimickers acted via the generation of H 2O 2 52<br />

and that H 2O 2 stimulates lipid synthesis in adipocytes. 53 In complementary studies,<br />

insulin was also shown to stimulate the generation of H 2O 2 in adipocytes. 54<br />

An early characterization of the enzymology of this process revealed that<br />

insulin activated a plasma membrane enzyme system with the properties of an<br />

NADPH oxidase, resulting in the downstream production of H 2O 2. 55,56 Further<br />

biochemical studies of this activity showed that it accounted for insulin-stimulated<br />

ROS production in rat adipocyte plasma membranes 57,58 and was also present in<br />

3T3-L1 adipocytes. 59 NADPH oxidase catalyzes the reduction of oxygen to superoxide<br />

radical: 2 O 2 + NADPH → 2 O 2 •– + NADP + + H + . While superoxide anions<br />

can react with thiols, they are rapidly converted spontaneously or by superoxide<br />

dismutase in the cell to generate H 2O 2. 42


182 Oxidative Stress and Inflammatory Mechanisms<br />

INSULIN-STIMULATED H 2O 2 GENERATION NEGATIVELY<br />

REGULATES PTP1B<br />

One hypothesis drawn from the diverse research data described above suggests<br />

that plasma membrane oxidase activity stimulated by insulin generates cellular<br />

ROS which, in turn, facilitates the insulin signaling cascade via the oxidative<br />

inhibition of cellular PTP activity, in particular involving PTP1B. 5,6 We reported<br />

that in the 3T3-L1 adipocyte cell model, insulin stimulated the generation of<br />

cellular ROS in minutes within the high physiologic range of insulin concentrations.<br />

30,31<br />

Blocking insulin-stimulated ROS with diphenyleneiodonium (DPI), an inhibitor<br />

of cellular NADPH oxidase activity, or catalase, reduced the insulin-stimulated<br />

autophosphorylation of the insulin receptor and the IRS proteins consistent<br />

with the notion that the oxidant signal inhibited cellular PTPs that serve as<br />

negative regulators of the insulin signaling cascade. That the enhancement of<br />

insulin signaling by the oxidant signal was associated with PTP inhibition was<br />

also shown by a novel approach that includes sample handling and analysis under<br />

anaerobic conditions to preserve the endogenous activity of PTPs isolated from<br />

cultured cells and avoiding cysteine oxidation that occurs on exposure to air. 60<br />

In HepG2 hepatoma cells, stimulation with 100 nM insulin for 5 min reduced<br />

overall PTP activity in the cell homogenate, and biochemical reduction of the<br />

enzyme samples with dithiothreitol prior to PTP assay fully restored the reduced<br />

PTP activity of the insulin-treated samples, indicating that they had been reversibly<br />

oxidized and inactivated by insulin exposure. 30<br />

Similar effects were observed in 3T3-L1 adipocytes. Insulin-stimulated generation<br />

of H 2O 2 also affected the specific activity of endogenous PTP1B isolated<br />

from intact cells. In 3T3-L1 adipocytes, insulin treatment also potently reduced<br />

the activity of immunoprecipitated PTP1B that was also reversible toward control<br />

levels by preincubation with dithiothreitol prior to assay. 30 In the continued<br />

presence of insulin, this effect was sustained for at least 10 min. Catalase pretreatment<br />

of the cells abolished the insulin-induced inhibition of PTP1B. Oxidative<br />

inactivation of PTP1B is thus associated with enhanced insulin signal transduction<br />

via the insulin-stimulated H 2O 2 signal.<br />

Further support of the notion of ROS serving as an enhancer of insulin action<br />

was provided by McClung and colleagues 61 who generated a line of mice overexpressing<br />

cellular glutathione peroxidase and showed that they developed hyperinsulinemia<br />

and hyperglycemia with insulin resistance and obesity. Insulin signaling<br />

was diminished in the tissues with reduced insulin-stimulated insulin<br />

receptor β-subunit phosphorylation and activation of Akt. Since excess glutathione<br />

peroxidase quenches intracellular ROS generation, it was hypothesized<br />

to interfere with insulin action by blocking PTP inactivation. Another group has<br />

also reported that an inability to generate ROS in response to insulin stimulation<br />

accounts for insulin insensitivity of ERK activation in SK-N-BE(2) neuroblastoma<br />

cells. 62


Oxygen Species and Signal Transduction 183<br />

DYNAMICS OF RECEPTOR-INDUCED ROS, PTP INHIBITION,<br />

<strong>AND</strong> SIGNAL TRANSDUCTION<br />

ROS production following ligand stimulation, including insulin, generates only<br />

a fraction of the ROS concentration observed in phagocytic cells and follows a<br />

brief time course, on the order of minutes. 30 These features apparently account<br />

for the signaling role of insulin-induced ROS compared to the chronic exposure<br />

to ROS in patients with hyperglycemia that is associated with organ dysfunction<br />

and chronic complications of diabetes mellitus. 63<br />

The low levels of ROS in insulin signaling imply specific cellular protein<br />

targets that are particularly susceptible to oxidative modification. This biochemical<br />

evidence, therefore, also supports the notion of a discrete network of “redox<br />

circuitry” 42,64,65 with temporal and spatial influences that are likely to correspond<br />

to other regulatory aspects of the insulin action pathway. 66 Work by Stone and<br />

colleagues 36 clearly describes the gradient of cellular responses to oxidative<br />

stimulation, with low levels eliciting signaling responses and cell proliferation<br />

without cellular oxidation of most free thiols such as glutathione. Increasing<br />

exposure to ROS is followed by growth arrest, cellular damage, thiol oxidation,<br />

and induced cell death.<br />

Insight into the disposition of growth factor-induced ROS has recently been<br />

gleaned from elegant studies by Reynolds et al. who modeled EGF receptor<br />

activation and signal propagation with PTP inhibition by ROS. 67 ROS generated<br />

in response to EGF stimulation in MCF7 cells were spatially constrained to a<br />

layer below the plasma membrane.<br />

Using reaction constants gleaned from published experimental work, including<br />

PTP inhibition by H 2O 2 and related effects, a model was developed that was<br />

most consistent with a bistable activation state for PTP and receptor tyrosine<br />

kinase activity. The formation of a reaction “wavefront” is postulated to involve<br />

local cycles of EGFR activation, H 2O 2 production, and PTP inhibition, which<br />

propagates along the plasma membrane. The model proposes that signal initiation<br />

involving oxidative inhibition of PTPs adds a feedback control loop to a reaction<br />

network that responds in an amplified and switch-like manner, especially at low<br />

levels of ligand stimulus.<br />

Consistent with this model, they also showed experimentally that blocking<br />

ligand-dependent H 2O 2 generation with the NADPH oxidase inhibitor DPI abolished<br />

the propagation of receptor phosphorylation and the amplification of receptor<br />

activation at low concentrations of EGF, converting the system to a “stable”<br />

steady state by generating a more linear phosphorylation response to ligand<br />

stimulation. 67 Diminishing PTP inactivation with DPI also suppresses insulin<br />

receptor activation and several aspects of the downstream insulin signaling cascade.<br />

31,68 Thus, it would be of interest to employ these types of novel imaging<br />

techniques to evaluate the role of similar regulatory networks in insulin-sensitive<br />

cells.


184 Oxidative Stress and Inflammatory Mechanisms<br />

IDENTIFICATION OF THE NADPH OXIDASE HOMOLOG NOX4<br />

AS A POTENTIAL MEDIATOR OF INSULIN-STIMULATED ROS<br />

Using the prototypic NADPH oxidase (Nox) catalytic gp91phox subunit, Cheng<br />

and Lambeth cloned a small family of five homologous Nox catalytic subunits<br />

(reviewed in Lambeth 69 ). Although Nox4 is prominently expressed in kidney, 70–72<br />

we also showed that it was expressed among insulin-sensitive cell types including<br />

liver, skeletal muscles, and adipocytes. 4 Evidence for an integral role of Nox4 in<br />

the insulin-induced oxidant signal was obtained by adenovirus-mediated expression<br />

of Nox4 deletion constructs lacking either the NADPH binding domain or<br />

the combined FAD/NADPH domains that acted in a dominant-negative fashion<br />

and attenuated insulin-stimulated generation of H 2O 2.<br />

Functionally, expression of the deletion constructs led to an inhibition of<br />

insulin receptor and IRS-1 tyrosine phosphorylation, activation of downstream<br />

serine kinases, and glucose uptake. Similar results were obtained with transfection<br />

of siRNA oligonucleotides that reduced Nox4 protein abundance. 4 Altogether,<br />

these results suggest that Nox4 overexpression potentiates, and reduced Nox4<br />

mass diminishes, insulin signal transduction in 3T3-L1 cells.<br />

ROS generation by Nox4 in adipocytes was also associated with oxidative<br />

inhibition of cellular PTP1B activity. 4 Overexpression of recombinant PTP1B<br />

inhibited insulin-stimulated tyrosine phosphorylation of the insulin receptor,<br />

which was significantly reversed by co-overexpression of active Nox4. The effect<br />

of overexpression of Nox4 on receptor autophosphorylation was closely associated<br />

with inhibition of PTP1B catalytic activity, measured in enzyme immunoprecipitates.<br />

REGULATION OF NOX4 SIGNALING IN INSULIN<br />

ACTION CASCADE<br />

Clarification of the regulation of Nox4 by growth factors has been elusive. 73 All<br />

the Nox catalytic subunits analogous to gp91phox have been shown to be physically<br />

associated with p22phox, an essential component of the flavocytochrome<br />

complex 74 ; thus, regulation of p22 abundance may be a means of regulating Nox4<br />

activity. The phagocyte NADPH oxidase has been extensively characterized; the<br />

regulatory proteins p47phox and p67phox have been shown to play a key role in<br />

protein complex formation with activation of Nox2. 69,75<br />

Since these proteins are not expressed in non-phagocytic cells, several groups<br />

have recently reported the cloning of related interacting proteins termed NOXO1<br />

(Nox organizer 1 or p41) and NOXA1 (Nox activator 1 or p51) that are homologs<br />

of the phagocyte p47phox and p67phox proteins. 76–79 Superoxide production by<br />

Nox1 and Nox3 is enhanced by interactions with these regulatory subunits. 80–83<br />

Nox5 is regulated by intracellular calcium levels. 84 NOXO1 and NOXA1 do not<br />

affect Nox4 activity and currently the mechanisms underlying the regulation of<br />

Nox4 activity are not known. NIH 3T3 fibroblasts transfected with Nox4 exhibit


Oxygen Species and Signal Transduction 185<br />

constitutively increased superoxide generation, 70 suggesting that Nox4 may not<br />

depend on activation by regulatory subunits. However, the mechanism of the rapid<br />

activation of Nox4 in insulin-stimulated ROS generation remains unexplained.<br />

POTENTIAL ROLE OF RAC IN INSULIN-STIMULATED H 2O 2<br />

<strong>AND</strong> PTP REGULATION<br />

Rac is a key component of the NADPH oxidase complex in a variety of cell<br />

types. 73,85 A chimeric Rac1-p67phox protein increases Nox2 activity, expression<br />

of a dominant-negative Rac inhibits the rise in ROS seen after stimulation by<br />

growth factors or cytokines, and a constitutively active Rac stimulates ROS<br />

formation in NIH 3T3 cells and in renal mesangial cells. 86,87 Since superoxide<br />

generation in the renal cell system is likely to involve Nox4, these data also<br />

implicated Rac as a potential regulator of this Nox homolog. 87<br />

In insulin-sensitive cells, Rac is involved in distal insulin signaling to glucose<br />

transport. 88 However, to date, our experiments using overexpression of wild-type,<br />

dominant-negative, and constitutively active Rac in differentiated 3T3-L1 adipocytes<br />

have not demonstrated a role for Rac in insulin-stimulated receptor autophosphorylation<br />

or substrate (IRS-1) tyrosine phosphorylation.<br />

GαI2<br />

Data from a variety of sources over the years have also linked the small G-protein<br />

Gαi2 with insulin action and potentially with insulin-stimulated NADPH oxidase<br />

activity (reviewed by Waters et al. and Malbon 89,90 ). Insulin-stimulated plasma<br />

membrane NADPH oxidase was shown to be coupled to Gαi2 91 and more recently<br />

insulin stimulation led to protein association between Gαi2 and the insulin receptor.<br />

92 Insulin receptor autophosphorylation was stimulated by activated Gαi2, and<br />

blocked by pretreatment with pertussis toxin, consistent with an earlier paper on<br />

Fao hepatoma cells. 93<br />

A recent paper also linked the attenuation of platelet activation by insulin<br />

with tyrosine phosphorylation of Gαi2 and complex formation between IRS-1<br />

and Gαi2, but not other Gα subunits. 94 Other approaches including a series of<br />

studies by Malbon and colleagues in transgenic mice have supported a permissive<br />

role of Gαi2 in insulin signaling in vivo. 95–98 Interestingly, mice lacking Gαi2<br />

also exhibited increased tissue PTP activity, 95 implicating a potential loss of<br />

insulin-stimulated NADPH oxidase activity in Gαi2-deficient animals with<br />

reduced oxidative inhibition of PTPs that regulate the insulin action pathway.<br />

Overall, these studies are consistent with the hypothesis that the regulation of<br />

tyrosine phosphorylation in the insulin signal cascade is propagated by a wave<br />

of H 2O 2, possibly generated by a link between the insulin receptor and Gαi2,<br />

coupled to cellular NADPH oxidase activity, which transiently inhibits PTP<br />

activities.


186 Oxidative Stress and Inflammatory Mechanisms<br />

NOVEL TARGETS OF INSULIN-STIMULATED ROS THAT MAY<br />

POTENTIALLY INFLUENCE INSULIN ACTION<br />

In addition to PTPs that have been implicated in the regulation of insulin signaling<br />

(e.g., PTP1B and LAR, as discussed above 11 ), a number of additional cellular<br />

enzymes are potential targets of oxidative inhibition by insulin-induced ROS. The<br />

serine–threonine phosphatase PP2A implicated in the negative regulation of Akt<br />

by dephosphorylation of ser-473 has a redox-sensitive cysteine residue that is<br />

potentially susceptible to inhibition by H 2O 2. 99<br />

The dual-specificity (ser/thr and tyr) phosphatase MKP-1, which attenuates<br />

insulin-stimulated MAP kinase activity, 100 is also dependent on a reduced thiol<br />

for activity. The lipid phosphatase PTEN can modulate downstream insulin signaling,<br />

101 and is also inactivated by oxidation of essential cysteine residues in its<br />

active site, which can be reactivated by thioredoxin in cells. 102,103 Further research<br />

is needed to determine the effects of the oxidative inhibition of these important<br />

signaling regulators on proximal and distal events in the insulin action cascade.<br />

Rhee and his group have shown that even in a cellular milieu containing<br />

millimolar concentrations of slowly reactive thiols like GSH, only a limited set<br />

of proteins are rapidly oxidized by growth factor-stimulated ROS, including<br />

PTP1B and a few other proteins with reactive cysteines including protein disulfide<br />

isomerases, thioredoxin reductase, and creatine kinase. 104,105 We have confirmed<br />

similar findings following cellular insulin stimulation of insulin target cell types<br />

(Wu et al., unpublished data), and are currently employing novel fluorescently<br />

tagged thiol reagents (iodoacetamide and maleimide derivatives) and adapting<br />

proteomic methods for the differential labeling of protein thiol before and after<br />

cellular insulin stimulation.<br />

Further work will help define the regulatory components and mechanism of<br />

NADPH oxidase activation by insulin in various insulin-sensitive cell types and<br />

the effects of insulin-stimulated ROS on the insulin action cascade with the<br />

identification of specific cellular targets susceptible to oxidative modification by<br />

insulin-stimulated ROS. Elucidation of these processes will determine how they<br />

are involved in the normal physiology of insulin signaling, whether they contribute<br />

to insulin-resistant disease states, and whether elements of this system may<br />

emerge as novel targets for pharmaceutical intervention.<br />

SUMMARY<br />

Reversible tyrosine phosphorylation plays an essential role in the regulation of<br />

transmission of the insulin signal at receptor and post-receptor sites in the insulin<br />

action pathway. PTPs, in particular PTP1B, and other thiol-sensitive signaling<br />

proteins are integral to the negative regulation of insulin signaling. A growing<br />

body of data over the past three decades has led to the appreciation that cellular<br />

stimulation with insulin generates ROS that can inhibit these negative regulators<br />

by oxidative biochemical alterations which, in turn, can facilitate the insulin<br />

signaling cascade. With the recognition that a small family of NADPH oxidase


Oxygen Species and Signal Transduction 187<br />

α-Subunit<br />

β-Subunit<br />

FIGURE 11.2 Insulin-induced ROS production and PTP regulation via Nox4 in insulinsensitive<br />

cells. This figure illustrates the action of insulin to stimulate receptor tyrosine<br />

autophosphorylation which activates the receptor toward its cellular substrate (IRS) proteins.<br />

The receptor and IRS tyrosine phosphorylation require cellular PTP activity to return<br />

to a basal state. The insulin receptor is coupled to the Nox4 homolog of NADPH oxidase<br />

and stimulates the cellular generation of reactive oxygen species which, in turn, leads to<br />

oxidative inhibition of the thiol-dependent PTPs, including PTP1B, the major phosphatase<br />

for the insulin signaling cascade. The lower right portion illustrates the PTP reaction<br />

mechanism in which the reduced thiol side chain in the enzyme catalytic domain forms<br />

a phosphocysteine intermediate with the phosphotyrosine substrate. If the catalytic PTP<br />

thiol is oxidized, this reaction intermediate cannot form and the enzymatic reaction is<br />

blocked. See text for discussion and references.<br />

homologs catalyzes the generation of ROS at the plasma membrane, we have<br />

recently provided evidence in the 3T3-L1 adipocyte system that Nox4, which is<br />

expressed in insulin-sensitive cell types, is a novel molecular target that may<br />

mediate this process (Figure 11.2).<br />

ACKNOWLEDGMENT<br />

This work was supported by National Institutes of Health grant RO1 DK43396<br />

to Dr. Goldstein.<br />

REFERENCES<br />

Insulin<br />

ATP<br />

-pY<br />

-pY<br />

-pY<br />

Y - Substrate Substrate - pY<br />

-PO 4<br />

PTP1B<br />

-PO 4<br />

PTP Cys<br />

215<br />

PTP thiol<br />

NADPH<br />

Oxidase<br />

(Nox4)<br />

pY protein<br />

1. Evans, J.L. et al. Oxidative stress and stress-activated signaling pathways: a<br />

unifying hypothesis of type 2 diabetes. Endocr. Rev. 23, 599, 2002.<br />

2. Ceriello, A. and Motz, E. Is oxidative stress the pathogenic mechanism underlying<br />

insulin resistance, diabetes, and cardiovascular disease? The common soil hypothesis<br />

revisited. Arterioscler. Thromb. Vasc. Biol. 24, 816, 2004.<br />

3. Houstis, N., Rosen, E.D., and Lander, E.S. Reactive oxygen species have a causal<br />

role in multiple forms of insulin resistance. Nature 440, 944, 2006.<br />

FAD<br />

+ NADP H+<br />

S −<br />

O<br />

−O−O P<br />

O


188 Oxidative Stress and Inflammatory Mechanisms<br />

4. Mahadev, K. et al. The NAD(P)H oxidase homolog Nox4 modulates insulinstimulated<br />

generation of H2O2 and plays an integral role in insulin signal transduction.<br />

Mol. Cell. Biol. 24, 1844, 2004.<br />

5. Goldstein, B.J., Mahadev, K., and Wu, X. Redox paradox: insulin action is facilitated<br />

by insulin-stimulated reactive oxygen species with multiple potential signaling<br />

targets. Diabetes 54, 311, 2005.<br />

6. Goldstein, B.J. et al. Role of insulin-induced reactive oxygen species in the insulin<br />

signaling pathway. Antioxid. Redox Signal. 7, 1021, 2005.<br />

7. Goldstein, B.J. Insulin resistance as the core defect in type 2 diabetes mellitus.<br />

Am. J. Cardiol. 90, 3, 2002.<br />

8. Stumvoll, M., Goldstein, B.J., and van Haeften, T.W. Type 2 diabetes: principles<br />

of pathogenesis and therapy. Lancet 365, 1333, 2005.<br />

9. Kido, Y., Nakae, J., and Accili, D. The insulin receptor and its cellular targets. J.<br />

Clin. Endocr. Metab. 86, 972, 2001.<br />

10. White, M.F. IRS proteins and the common path to diabetes. Am. J. Physiol.<br />

Endocrinol. Metab. 283, E413, 2002.<br />

11. Goldstein, B.J. Protein-tyrosine phosphatases and the regulation of insulin action,<br />

in Diabetes Mellitus: A Fundamental and Clinical Text, 3rd ed., LeRoith, D. et<br />

al., Eds., Lippincott, Philadelphia, 2003, p. 255.<br />

12. Goldstein, B.J. Protein-tyrosine phosphatase 1B (PTP1B): a novel therapeutic<br />

target for type 2 diabetes mellitus, obesity and related states of insulin resistance.<br />

Curr. Drug Targets Immune. Endocr. Metabol. Disord. 1, 265, 2001.<br />

13. Asante-Appiah, E. and Kennedy, B.P. Protein tyrosine phosphatases: the quest for<br />

negative regulators of insulin action. Am. J. Physiol. Endocrinol. Metab. 284,<br />

E663, 2003.<br />

14. Tonks, N.K. PTP1B: from the sidelines to the front lines! FEBS Lett. 546, 140,<br />

2003.<br />

15. Elchebly, M., Cheng, A., and Tremblay, M.L. Modulation of insulin signaling by<br />

protein tyrosine phosphatases. J. Molec. Med. 78, 473, 2000.<br />

16. Klaman, L.D. et al. Increased energy expenditure, decreased adiposity, and tissuespecific<br />

insulin sensitivity in protein-tyrosine phosphatase 1B-deficient mice. Mol.<br />

Cell. Biol. 20, 5479, 2000.<br />

17. Hashimoto, N., Zhang, W.R., and Goldstein, B.J. Insulin receptor and epidermal<br />

growth factor receptor dephosphorylation by three major rat liver protein-tyrosine<br />

phosphatases expressed in a recombinant bacterial system. Biochem. J. 284, 569,<br />

1992.<br />

18. Walchli, S. et al. Identification of tyrosine phosphatases that dephosphorylate the<br />

insulin receptor: a brute force approach based on “substrate-trapping” mutants. J.<br />

Biol. Chem. 275, 9792, 2000.<br />

19. Goldstein, B.J. et al. Tyrosine dephosphorylation and deactivation of insulin receptor<br />

substrate-1 by protein-tyrosine phosphatase 1B: possible facilitation by the<br />

formation of a ternary complex with the Grb2 adaptor protein. J. Biol. Chem. 275,<br />

4283, 2000.<br />

20. Calera, M.R., Vallega, G., and Pilch, P.F. Dynamics of protein–tyrosine phosphatases<br />

in rat adipocytes. J. Biol. Chem. 275, 6308, 2000.<br />

21. Puius, Y.A. et al. Identification of a second aryl phosphate-binding site in proteintyrosine<br />

phosphatase 1b: a paradigm for inhibitor design. Proc. Natl. Acad. Sci.<br />

USA 94, 13420, 1997.


Oxygen Species and Signal Transduction 189<br />

22. Salmeen, A. et al. Molecular basis for the dephosphorylation of the activation<br />

segment of the insulin receptor by protein tyrosine phosphatase 1B. Mol. Cell 6,<br />

1401, 2000.<br />

23. Dadke, S. and Chernoff, J. Interaction of protein tyrosine phosphatase (PTP) 1B<br />

with its substrates is influenced by two distinct binding domains. Biochem. J. 364,<br />

377, 2002.<br />

24. Tonks, N.K. and Neel, B.G. Combinatorial control of the specificity of protein<br />

tyrosine phosphatases. Curr. Opin. Cell Biol. 13, 182, 2001.<br />

25. Andersen, J.N. et al. Structural and evolutionary relationships among protein<br />

tyrosine phosphatase domains. Mol. Cell Biol. 21, 7117, 2001.<br />

26. Keyse, S.M. Protein phosphatases and the regulation of mitogen-activated protein<br />

kinase signalling. Curr. Opin. Cell Biol. 12, 186, 2000.<br />

27. Zhang, Z.Y. Protein tyrosine phosphatases: structure and function, substrate specificity,<br />

and inhibitor development. Annu. Rev. Pharmacol. Toxicol. 42, 209, 2002.<br />

28. Denu, J.M. and Dixon, J.E. Protein tyrosine phosphatases: mechanisms of catalysis<br />

and regulation. Curr. Opin. Chem. Biol. 2, 633, 1998.<br />

29. Zhang, Z.Y. Mechanistic studies on protein tyrosine phosphatases. Progr. Nucl.<br />

Acid Res. Mol. Biol. 73, 171, 2003.<br />

30. Mahadev, K. et al. Insulin-stimulated hydrogen peroxide reversibly inhibits protein-tyrosine<br />

phosphatase 1B in vivo and enhances the early insulin action cascade.<br />

J. Biol. Chem. 276, 21938, 2001.<br />

31. Mahadev, K. et al. Hydrogen peroxide generated during cellular insulin stimulation<br />

is integral to activation of the distal insulin signaling cascade in 3T3-L1 adipocytes.<br />

J. Biol. Chem. 276, 48662, 2001.<br />

32. Meng, T.C., Fukada, T., and Tonks, N.K. Reversible oxidation and inactivation of<br />

protein tyrosine phosphatases in vivo. Mol. Cell 9, 387, 2002.<br />

33. Rhee, S.G. et al. Intracellular messenger function of hydrogen peroxide and its<br />

regulation by peroxiredoxins. Curr. Opin. Cell Biol. 17, 183, 2005.<br />

34. Sundaresan, M. et al. Requirement for generation of H 2O 2 for platelet-derived<br />

growth factor signal transduction. Science 270, 296, 1995.<br />

35. Bae, Y.S. et al. Epidermal growth factor (EGF)-induced generation of hydrogen<br />

peroxide: role in EGF receptor-mediated tyrosine phosphorylation. J. Biol. Chem.<br />

272, 217, 1997.<br />

36. Stone, J.R. and Yang, S. Hydrogen peroxide: a signaling messenger. Antioxid.<br />

Redox Signal. 8, 243, 2006.<br />

37. Lee, S.R. et al. Reversible inactivation of protein-tyrosine phosphatase 1b in a431<br />

cells stimulated with epidermal growth factor. J. Biol. Chem. 273, 15366, 1998.<br />

38. DeGnore, J.P. et al. Identification of the oxidation states of the active site cysteine<br />

in a recombinant protein tyrosine phosphatase by electrospray mass spectrometry<br />

using on-line desalting. Rapid Commun. Mass Spectrom. 12, 1457, 1998.<br />

39. Meng, T.C. et al. Regulation of insulin signaling through reversible oxidation of<br />

the protein–tyrosine phosphatases TC45 and PTP1B. J. Biol. Chem. 279, 37716,<br />

2004.<br />

40. Denu, J.M. and Tanner, K.G. Specific and reversible inactivation of protein<br />

tyrosine phosphatases by hydrogen peroxide: evidence for a sulfenic acid intermediate<br />

and implications for redox regulation. Biochemistry 37, 5633, 1998.<br />

41. Barrett, W.C. et al. Roles of superoxide radical anion in signal transduction<br />

mediated by reversible regulation of protein-tyrosine phosphatase 1B. J. Biol.<br />

Chem. 274, 34543, 1999.


190 Oxidative Stress and Inflammatory Mechanisms<br />

42. Rhee, S.G. et al. Cellular regulation by hydrogen peroxide. J. Am. Soc. Nephrol.<br />

14, S211, 2003.<br />

43. Claiborne, A. et al. Protein sulfenic acids: diverse roles for an unlikely player in<br />

enzyme catalysis and redox regulation. Biochemistry 38, 15407, 1999.<br />

44. Claiborne, A. et al. Protein sulfenic acid stabilization and function in enzyme<br />

catalysis and gene regulation. FASEB J. 7, 1483, 1993.<br />

45. Skorey, K. et al. How does alendronate inhibit protein-tyrosine phosphatases. J.<br />

Biol. Chem. 272, 22472, 1997.<br />

46. Barrett, W.C. et al. Regulation of PTP1B via glutathionylation of the active site<br />

cysteine 215. Biochemistry 38, 6699, 1999.<br />

47. van Montfort, R.L. et al. Oxidation state of the active-site cysteine in protein<br />

tyrosine phosphatase 1B. Nature 423, 773, 2003.<br />

48. Salmeen, A. et al. Redox regulation of protein tyrosine phosphatase 1B involves<br />

a sulphenyl-amide intermediate. Nature 423, 769, 2003.<br />

49. Czech, M.P. and Fain, J.N. Cu ++ -dependent thiol stimulation of glucose metabolism<br />

in white fat cells. J. Biol. Chem. 247, 6218, 1972.<br />

50. Czech, M.P., Lawrence, J.C., Jr., and Lynn, W.S. Evidence for electron transfer<br />

reactions involved in the Cu 2+ -dependent thiol activation of fat cell glucose utilization.<br />

J. Biol. Chem. 249, 1001, 1974.<br />

51. May, J.M. The insulin-like effects of low molecular weight thiols: role of trace<br />

metal contamination of commercial thiols. Horm. Metab Res. 12, 587, 1980.<br />

52. Livingston, J.N., Gurny, P.A., and Lockwood, D.H. Insulin-like effects of<br />

polyamines in fat cells: mediation by H 2O 2 formation. J. Biol. Chem. 252, 560,<br />

1977.<br />

53. May, J.M. and de Haen, C. The insulin-like effect of hydrogen peroxide on<br />

pathways of lipid synthesis in rat adipocytes. J. Biol. Chem. 254, 9017, 1979.<br />

54. May, J.M. and de Haen, C. Insulin-stimulated intracellular hydrogen peroxide<br />

production in rat epididymal fat cells. J. Biol. Chem. 254, 2214, 1979.<br />

55. Mukherjee, S.P. and Lynn, W.S. Reduced nicotinamide adenine dinucleotide phosphate<br />

oxidase in adipocyte plasma membrane and its activation by insulin: possible<br />

role in the hormone’s effects on adenylate cyclase and the hexose monophosphate<br />

shunt. Arch. Biochem. Biophys. 184, 69, 1977.<br />

56. Mukherjee, S.P., Lane, R.H., and Lynn, W.S. Endogenous hydrogen peroxide and<br />

peroxidative metabolism in adipocytes in response to insulin and sulfhydryl<br />

reagents. Biochem. Pharmacol. 27, 2589, 1978.<br />

57. Krieger-Brauer, H.I. and Kather, H. Human fat cells possess a plasma membranebound<br />

H 2O 2-generating system that is activated by insulin via a mechanism<br />

bypassing the receptor kinase. J. Clin Invest. 89, 1006, 1992.<br />

58. Krieger-Brauer, H.I. and Kather, H. The stimulus-sensitive H 2O 2-generating system<br />

present in human fat-cell plasma membranes is multireceptor-linked and under<br />

antagonistic control by hormones and cytokines. Biochem. J. 307, 543, 1995.<br />

59. Krieger-Brauer, H.I. and Kather, H. Antagonistic effects of different members of<br />

the fibroblast and platelet-derived growth factor families on adipose conversion<br />

and NADPH-dependent H 2O 2 generation in 3T3 L1-cells. Biochem. J. 307, 549,<br />

1995.<br />

60. Zhu, L. et al. Use of an anaerobic environment to preserve the endogenous activity<br />

of protein-tyrosine phosphatases isolated from intact cells. FASEB J. 15, 1637,<br />

2001.


Oxygen Species and Signal Transduction 191<br />

61. McClung, J.P. et al. Development of insulin resistance and obesity in mice overexpressing<br />

cellular glutathione peroxidase. Proc. Natl. Acad. Sci. USA 101, 8852,<br />

2004.<br />

62. Hwang, J.J. and Hur, K.C. Insulin cannot activate extracellular-signal-related<br />

kinase due to inability to generate reactive oxygen species in SK-N-BE(2) human<br />

neuroblastoma cells. Mol. Cell 20, 280, 2005.<br />

63. Brownlee, M. Biochemistry and molecular cell biology of diabetic complications.<br />

Nature 414, 813, 2001.<br />

64. Go, Y.M. et al. H 2O 2-dependent activation of GCLC-ARE4 reporter occurs by<br />

mitogen-activated protein kinase pathways without oxidation of cellular glutathione<br />

or thioredoxin-1. J. Biol. Chem. 279, 5837, 2004.<br />

65. Dooley, C.T. et al. Imaging dynamic redox changes in mammalian cells with green<br />

fluorescent protein indicators. J. Biol. Chem. 279, 22284, 2004.<br />

66. Saltiel, A.R. and Pessin, J.E. Insulin signaling in microdomains of the plasma<br />

membrane. Traffic 4, 711, 2003.<br />

67. Reynolds, A.R. et al. EGFR activation coupled to inhibition of tyrosine phosphatases<br />

causes lateral signal propagation. Nat. Cell. Biol. 5, 447, 2003.<br />

68. Mahadev, K. et al. Integration of multiple downstream signals determines the net<br />

effect of insulin on MAP kinase vs. PI 3-kinase activation: potential role of insulinstimulated<br />

H 2O 2. Cell Signal. 16, 323, 2004.<br />

69. Lambeth, J.D. NOX enzymes and the biology of reactive oxygen. Nat. Rev.<br />

Immunol. 4, 181, 2004.<br />

70. Geiszt, M. et al. Identification of renox, an NAD(P)H oxidase in kidney. Proc.<br />

Natl. Acad. Sci. USA 97, 8010, 2000.<br />

71. Cheng, G. et al. Homologs of gp91phox: cloning and tissue expression of Nox3,<br />

Nox4, and Nox5. Gene 269, 131, 2001.<br />

72. Shiose, A. et al. A novel superoxide-producing NAD(P)H oxidase in kidney. J.<br />

Biol. Chem. 276, 1417, 2001.<br />

73. Hordijk, P.L. Regulation of NADPH oxidases: role of Rac proteins. Circ. Res. 98,<br />

453, 2006.<br />

74. Ambasta, R.K., et al., Direct interaction of the novel Nox proteins with p22phox<br />

is required for the formation of a functionally active NADPH oxidase. J. Biol.<br />

Chem. 279, 45935, 2004.<br />

75. Babior, B.M., Lambeth, J.D., and Nauseef, W. The neutrophil NADPH oxidase.<br />

Arch. Biochem. Biophys. 397, 342, 2002.<br />

76. Banfi, B. et al. Two novel proteins activate superoxide generation by the NADPH<br />

oxidase NOX1. J. Biol. Chem. 278, 3510, 2003.<br />

77. Takeya, R. et al. Novel human homologues of p47phox and p67phox participate<br />

in activation of superoxide-producing NADPH oxidases. J. Biol. Chem. 278,<br />

25234, 2003.<br />

78. Geiszt, M. et al. Proteins homologous to p47phox and p67phox support superoxide<br />

production by NAD(P)H oxidase 1 in colon epithelial cells. J. Biol. Chem. 278,<br />

20006, 2003.<br />

79. Cheng, G. and Lambeth, J.D. NOXO1, regulation of lipid binding, localization,<br />

and activation of Nox1 by the Phox homology (PX) domain. J. Biol. Chem. 279,<br />

4737, 2004.<br />

80. Cheng, G., Ritsick, D., and Lambeth, J.D. Nox3 regulation by NOXO1, p47phox,<br />

and p67phox. J. Biol. Chem. 279, 34250, 2004.


192 Oxidative Stress and Inflammatory Mechanisms<br />

81. Banfi, B. et al. NOX3: a superoxide-generating NADPH oxidase of the inner ear.<br />

J. Biol. Chem. 279, 46065, 2004.<br />

82. Cheng, G. and Lambeth, J.D. Alternative mRNA splice forms of NOXO1: differential<br />

tissue expression and regulation of Nox1 and Nox3. Gene 356, 118, 2005.<br />

83. Cheng, G. et al. Nox1-dependent reactive oxygen generation is regulated by Rac1.<br />

J. Biol. Chem. epub M512751200, 2006.<br />

84. Banfi, B. et al. Mechanism of Ca 2+ activation of the NADPH oxidase 5 (NOX5).<br />

J. Biol. Chem. 279, 18583, 2004.<br />

85. Lassegue, B. and Clempus, R.E. Vascular NAD(P)H oxidases: specific features,<br />

expression, and regulation. Am. J. Physiol. Regul. Integr. Comp Physiol. 285,<br />

R277, 2003.<br />

86. Sundaresan, M. et al. Regulation of reactive-oxygen-species generation in fibroblasts<br />

by Rac1. Biochem. J. 318 (Pt. 2), 379, 1996.<br />

87. Gorin, Y. et al. Nox4 mediates angiotensin II-induced activation of Akt/protein<br />

kinase B in mesangial cells. Am. J. Physiol. Renal Physiol. 285, F219, 2003.<br />

88. JeBailey, L. et al. Skeletal muscle cells and adipocytes differ in their reliance on<br />

TC10 and Rac for insulin-induced actin remodeling. Mol. Endocrinol. 18, 359,<br />

2004.<br />

89. Waters, C., Pyne, S., and Pyne, N.J. The role of G-protein coupled receptors and<br />

associated proteins in receptor tyrosine kinase signal transduction. Semin. Cell<br />

Dev. Biol. 15, 309, 2004.<br />

90. Malbon, C.C. Insulin signalling: putting the ‘G-’ in protein–protein interactions.<br />

Biochem. J. 380, e11, 2004.<br />

91. Krieger-Brauer, H.I., Medda, P.K., and Kather, H. Insulin-induced activation of<br />

NADPH-dependent H 2O 2 generation in human adipocyte plasma membranes is<br />

mediated by Galphai2. J. Biol. Chem. 272, 10135, 1997.<br />

92. Kreuzer, J., Nurnberg, B., and Krieger-Brauer, H.I. Ligand-dependent auto-phosphorylation<br />

of the insulin receptor is positively regulated by Gi-proteins. Biochem.<br />

J. 380, 831, 2004.<br />

93. Muller-Wieland, D. et al. Pertussis toxin inhibits autophosphorylation and activation<br />

of the insulin receptor kinase. Biochem. Biophys. Res. Commun. 181, 1479,<br />

1991.<br />

94. Ferreira, I.A. et al. IRS-1 mediates inhibition of Ca 2+ mobilization by insulin via<br />

the inhibitory G-protein Gi. J. Biol. Chem. 279, 3254, 2004.<br />

95. Moxham, C.M. and Malbon, C.C. Insulin action impaired by deficiency of the Gprotein<br />

subunit Gi-alpha-2. Nature 379, 840, 1996.<br />

96. Chen, J.F. et al. Conditional, tissue-specific expression of Q205L G alpha i2 in<br />

vivo mimics insulin action. J. Mol. Med. 75, 283, 1997.<br />

97. Zheng, X.L. et al. Expression of constitutively activated Gi-alpha-2 in vivo ameliorates<br />

streptozotocin-induced diabetes. J. Biol. Chem. 273, 23649, 1998.<br />

98. Song, X. et al. G-alpha-i2 enhances in vivo activation of and insulin signaling to<br />

GLUT4. J. Biol. Chem. 276, 34651, 2001.<br />

99. Guy, G.R. et al. Inactivation of a redox-sensitive protein phosphatase during the<br />

early events of tumor necrosis factor/interleukin-1 signal transduction. J. Biol.<br />

Chem. 268, 2141, 1993.<br />

100. Kusari, A.B. et al. Insulin-induced mitogen-activated protein (MAP) kinase phosphatase-1<br />

(MKP-1) attenuates insulin-stimulated MAP kinase activity: a mechanism<br />

for the feedback inhibition of insulin signaling. Mol. Endocrinol. 11, 1532,<br />

1997.


Oxygen Species and Signal Transduction 193<br />

101. Nakashima, N. et al. The tumor suppressor PTEN negatively regulates insulin<br />

signaling in 3T3-L1 adipocytes. J. Biol. Chem. 275, 12889, 2000.<br />

102. Lee, S.R. et al. Reversible inactivation of the tumor suppressor PTEN by H2O2.<br />

J. Biol. Chem. 277, 20336, 2002.<br />

103. Leslie, N.R. et al. Redox regulation of PI 3-kinase signalling via inactivation of<br />

PTEN. EMBO J. 22, 5501, 2003.<br />

104. Wu, Y., Kwon, K.S., and Rhee, S.G. Probing cellular protein targets of H 2O 2 with<br />

fluorescein-conjugated iodoacetamide and antibodies to fluorescein. FEBS Lett.<br />

440, 111, 1998.<br />

105. Kim, J.R. et al. Identification of proteins containing cysteine residues that are<br />

sensitive to oxidation by hydrogen peroxide at neutral pH. Anal. Biochem. 283,<br />

214, 2000.<br />

106. Finkel, T. Oxidant signals and oxidative stress. Curr. Opin. Cell Biol. 15, 247,<br />

2003.


12<br />

CONTENTS<br />

Intracellular Signaling<br />

Pathways and<br />

Peroxisome Proliferator-<br />

Activated Receptors in<br />

Vascular Health in<br />

Hypertension and in<br />

Diabetes<br />

Farhad Amiri, Karim Benkirane, and<br />

Ernesto L. Schiffrin<br />

Introduction .......................................................................................................196<br />

MAPK Signaling in Vascular Remodeling.......................................................198<br />

ERK1 and ERK2 Signaling, Effects of PPAR-γ, and Vascular Remodeling ..199<br />

p38 Signaling in Vascular Remodeling ............................................................200<br />

Role of JNK Signaling in Vascular Remodeling..............................................201<br />

PI3K Signaling Pathway and Vascular Remodeling ........................................201<br />

Role of Akt/PKB Signaling Pathway in Vascular Remodeling .......................202<br />

ROS and Vascular Remodeling.........................................................................202<br />

Vascular Inflammation and Remodeling...........................................................203<br />

Dual PPAR Activators.......................................................................................204<br />

Conclusion.........................................................................................................204<br />

References .........................................................................................................205<br />

195


196 Oxidative Stress and Inflammatory Mechanisms<br />

INTRODUCTION<br />

In both hypertension and diabetes mellitus significant changes that occur in the<br />

vasculature affect both large and small arteries and lead to cardiovascular events<br />

such as myocardial infarction, stroke, peripheral vascular disease, and compromise<br />

of renal function. In addition, diabetes also involves microvascular disease<br />

in different vascular beds including the retina and the kidney that contribute to<br />

the morbidity, and in the case of the kidney, the mortality associated with diabetes.<br />

The degree and importance of vascular disease and its contribution to mortality<br />

in diabetes are such that diabetes has been called a vascular disease although<br />

its origin is undoubtedly metabolic. Moreover, hypertension and diabetes are<br />

often associated: approximately 20% of hypertensives may become diabetic and<br />

80 to 90% of type 2 diabetic patients develop hypertension. Description of the<br />

nature of vascular disease in hypertension and diabetes, its mechanisms, and<br />

therapeutic targets, potential and already demonstrated, become accordingly<br />

extremely important. Some of these will be dealt with in this chapter, particularly<br />

in relation to signaling pathways and putative vascular protective effects of<br />

activation of peroxisome proliferator-activated receptors (PPARs).<br />

Vascular injuries in large arteries in hypertension and diabetes differ mainly<br />

in intensity and are much more severe in diabetes. In large arteries, injuries of<br />

both the intima with development of atherosclerosis and of the media with<br />

arteriosclerosis occur. The description of the atherosclerotic process is beyond<br />

the scope of this chapter. It is sufficient to say that the severity of atherosclerosis<br />

in diabetes is such that it affects the peripheral circulation, leading to amputation<br />

of lower limbs, and also affects the coronary circulation where diabetes is considered<br />

a “coronary equivalent.” 1<br />

In small (resistance) arteries that measure 150 to 300 µm in lumen diameter,<br />

hypertension is associated with remodeling changes that lead to increased peripheral<br />

resistance — the hallmark of essential hypertension. The remodeling of small<br />

arteries in hypertension is eutrophic — reduced outer and lumen diameters with<br />

increased media-to-lumen ratio and no significant increase in cross-sectional area<br />

of the media. 2–4 Associated with the structure changes are deposition in the media<br />

of extracellular matrix components such as collagen and fibronectin 5 and dysfunction<br />

of the endothelium. 6 Diabetes, on the other hand, involves hypertrophic<br />

remodeling in which the media-to-lumen ratio is also increased along with an<br />

increased media cross-section, achieving true media hypertrophy. 7,8 Deposition<br />

of collagen is less important. More frequently than in hypertension, endothelial<br />

dysfunction is the norm. 8,9<br />

In the United Kingdom Prospective Diabetes Study (UKPDS), tight blood<br />

pressure (BP) controls in hypertensive patients with type 2 diabetes reduced the<br />

risk of macrovascular disease, stroke, and deaths related to diabetes. 10 Most<br />

hypertension randomized clinical trials failed to show beneficial effects on cardiac<br />

ischemia expected from population studies. Thus, blood pressure lowering may<br />

not be enough to normalize remodeled arteries in hypertensive or diabetic subjects.<br />

In hypertensive patients, the extent and consequences of tissue ischemia


Receptors in Vascular Health in Hypertension and Diabetes 197<br />

(in the heart, kidney, or brain) are influenced by small vessel disease. 3 The<br />

intermediate coronary lesions that are frequent in hypertensive and diabetic subjects<br />

will produce changes in flow if small artery remodeling is present.<br />

The renin–angiotensin–aldosterone system (RAAS) plays a critical role in<br />

the initiation and progression of cardiovascular disease. 11 The RAAS contributes<br />

to vascular remodeling of small arteries through different pathways. Angiotensin<br />

II (Ang II) can indirectly promote vascular remodeling through hemodynamic<br />

effects or directly activate myriad intracellular signaling pathways such as mitogen-activated<br />

protein kinase (MAPK), phosphoinositide-3 kinase (PI3K), Janus<br />

kinase/signal transducers and activators of transcription (JAK/STAT), and reactive<br />

oxygen species (ROS) through AT 1 receptors. Ang II may also transactivate<br />

growth factor receptors such as epidermal growth factor receptor (EGFR). 12 These<br />

pathways in turn initiate cascades in which different key proteins are stimulated,<br />

including nuclear factors that are responsible for cardiovascular gene transcription.<br />

The involvement of RAAS in hypertension and diabetes-related complications<br />

has been clarified through the use of a variety of RAAS inhibitors including<br />

but not limited to angiotensin converting enzyme (ACE) inhibitors, Ang receptor<br />

blockers (ARB), and aldosterone receptor antagonists. 3<br />

Insulin-sensitizing thiazolidinediones (TZDs) or glitazones may exert protective<br />

cardiovascular properties in part through the prevention of vascular remodeling.<br />

13 For instance, TZDs such as pioglitazone and rosiglitazone reduced blood<br />

pressure in several hypertensive rodent models including Ang II-infused rats,<br />

stroke-prone spontaneously hypertensive rats (SHRSP), and deoxycorticosterone<br />

acetate (DOCA)-salt hypertensive rats. 14–16 Blood pressure lowering was accompanied<br />

by prevention of vascular remodeling and endothelial dysfunction and<br />

down-regulation of inflammatory mediators in these rodent models. Although the<br />

hypotensive effects of TZDs observed in rodents were not observed in humans,<br />

other beneficial effects such as prevention of vascular remodeling, improvement<br />

of endothelial function, 17,18 and down-regulation of inflammatory markers 19 have<br />

been reported.<br />

TZDs mediate their effects through binding of PPAR-γ, a ligand-activated<br />

transcription factor belonging to the nuclear receptor superfamily. Forming a<br />

dimer with RXR and associated to co-activators and co-repressors, PPAR-γ binds<br />

to DNA PPAR response elements (PPREs) and regulates many genes implicated<br />

in carbohydrate metabolism, inflammation and thrombosis, endothelial function<br />

and cell growth. 20 Other PPAR isoforms include PPAR-α, which is activated by<br />

fatty acids and by lipid lowering fibrates and predominantly expressed in tissues<br />

exhibiting high fatty acid catabolism such as liver, heart, kidney, and skeletal<br />

muscle. Another isoform is PPAR-β/δ, which is expressed ubiquitously and is<br />

involved in fatty acid oxidation. 21<br />

The major beneficial effects of TZDs have been attributed to their actions on<br />

several protein kinases such has MAPK, PI3K, and ROS-generating enzymes, all<br />

of which have been implicated in vascular remodeling in hypertension. 14,22


198 Oxidative Stress and Inflammatory Mechanisms<br />

MAPK SIGNALING IN VASCULAR REMODELING<br />

MAPK signaling has been largely studied for processes such as hyperplasia and<br />

hypertrophy associated with cell growth, and pro-inflammatory pathways, all of<br />

which are found in hypertensive and diabetic vascular remodeling. These<br />

serine/threonine kinases are subdivided into six subfamilies: extracellular signalregulated<br />

kinases 1 and 2 (ERK1/2), p38, c-jun N-terminal protein kinase (JNK),<br />

ERK3, ERK5, and ERK6. 23 Due to elaborate cross-talk among these kinases, we<br />

will focus only on ERK1/2, p38 and JNK proteins because they are the most<br />

commonly studied (Figure 12.1).<br />

Activation of MAPK is associated with cell growth, programmed cell death<br />

(apoptosis), cell transformation and differentiation, and cell contractility. ERK1/2<br />

proteins can be activated by different growth factors whereas p38 and JNK<br />

proteins are generally activated by cytokines and cellular stress. 23 Ang II is a<br />

potent stimulator of MAPK pathways through various upstream second messenger<br />

systems. 12,24,25<br />

MAPK activation requires phosphorylation of threonine and tyrosine residues<br />

by other kinases such as mixed lineage kinases (MLKs), MAPK kinase, or<br />

mitogen extracellular-regulated kinase (MEK). MEKs are activated by<br />

serine/threonine MEK kinases (Raf-1, A-Raf and B-Raf), which are in turn<br />

activated by small protein G (Ras family) and other kinases. 26,27 Raf phosphorylation<br />

is influenced by different kinases such as c-Src, protein kinase C (PKC),<br />

protein kinase B (PKB), and p21 (rac/Cdc42)-activated protein kinase (PAK). 26<br />

Stress, cytokines,<br />

growth factors Growth factors<br />

MLKs MLKs Raf<br />

MEK3/6 MEK4/7 MEK1/2<br />

p38 JNK<br />

Apoptosis, cell growth and<br />

differentiation, inflammation<br />

ERK1/2<br />

Cell growth and<br />

differentiation<br />

PD98059<br />

Stimulus<br />

MEKK<br />

MEK<br />

MAPK<br />

Biological<br />

response<br />

FIGURE 12.1 Mitogen-activated protein kinase (MAPK) signaling pathways and their<br />

responses. MLK = mixed lineage kinase (generic MEK). MEK = MAPK kinase. MEKK<br />

= MAPK kinase kinase.


Receptors in Vascular Health in Hypertension and Diabetes 199<br />

ERK1 <strong>AND</strong> ERK2 SIGNALING, EFFECTS OF PPAR-γ, <strong>AND</strong><br />

VASCULAR REMODELING<br />

ERK1 and ERK2 are ubiquitous proteins highly expressed in different tissues of<br />

the cardiovascular system (blood vessels, kidneys, heart). 28 Additionally, differential<br />

regulation occurs in different cell types. ERK2 expression was significantly<br />

higher than ERK1 expression in immune cells, 26 but both enzymes are equally<br />

and highly expressed in endothelial cells and vascular smooth muscle cells<br />

(VSMCs). 29–31 Ang II through binding to AT 1 receptors activates signaling cascades<br />

such as the Src homology 2 domain (Shc), c-Src, growth factor receptorbound<br />

protein 2 (Grb2), Son of sevenless (Sos), Ras-GTP, and MEK1, which<br />

then activate ERK1/2 through phosphorylation. 32<br />

Several studies using the MEK specific inhibitor PD98059 demonstrated an<br />

important role of ERK1/2 in the development and the maintenance of hypertension.<br />

33,34 For instance, in hypertensive vascular remodeling, ERK1/2 activation<br />

led to activation of phospholipase A 2, cyclooxygenase (COX)-2, and p90 ribosomal<br />

S6 kinase (p90Rsk), along with translocation of nuclear receptors and reorganization<br />

of cytoskeletal proteins implicated in cell growth and migration. 26<br />

More specifically, once ERK1/2 is phosphorylated, it will translocate to the<br />

nucleus and activate transcription of cell cycle genes. 23,26<br />

In VSMCs, ERK1/2 phosphorylation induced p90Rsk activation leading to<br />

ribosomal S6 protein phosphorylation and protein synthesis. In VSMCs derived<br />

from mesenteric arteries, we demonstrated the stimulatory effect of Ang II on<br />

cell hypertrophy, proliferation, and contractility. 29 Ang II-induced ERK1/2 activity<br />

was inhibited by PPAR-γ activation in conduit vessels whereas no effect was<br />

observed in mesenteric resistance vessels. 14 Unlike results of in vivo studies, in<br />

vitro acute stimulation of mesenteric VSMCs with Ang II-induced ERK1/2 activation<br />

was abrogated by PPAR-γ pre-stimulation with rosiglitazone (Figure<br />

12.2). 14 Taken together, these results suggest that the extracellular environment<br />

and duration of stimulation (acute versus chronic) affect PPAR-γ-modulated<br />

changes in Ang II-induced ERK1/2 activation.<br />

ERK1/2 may on the other hand inhibit PPAR-γ activity. In vivo, insulin, which<br />

constitutively activates MEKK1, induced in a ligand-independent manner PPAR-γ<br />

phosphorylation which increased TZD-dependent PPAR-γ trans-activation properties.<br />

35 In contrast, both epidermal growth factor (EGF) and platelet derived<br />

growth factor (PDGF) reduced PPAR-γ transcriptional activity in a ligand-dependent<br />

manner in adipocytes contained in the vascular periadventitial fat that regulates<br />

vascular function in paracrine fashion and may play an important role in<br />

blood pressure regulation and vascular remodeling. 36<br />

Although mesenteric artery PPAR-γ activity was reduced in Ang II-infused<br />

rats, albeit in the absence of changes in PPAR-γ expression, treatment with<br />

rosiglitazone prevented these changes, suggesting a potent negative regulatory<br />

role of PPAR-γ activators. 14 Endogenous PPAR ligands such as linoleic and<br />

retinoic acid may stimulate ERK1/2 activity in adipocytes and aortic VSMCs,<br />

respectively, 37,38 whereas prostaglandin J2 (PGJ2) activated ERK1/2 through PI3K


200 Oxidative Stress and Inflammatory Mechanisms<br />

Resistance Artery SMC Resistance<br />

Artery<br />

PD98059<br />

Ang II/AT 1<br />

MEK<br />

ERK1/2<br />

Transcription<br />

factors<br />

Cell growth<br />

Ang II/AT 1<br />

No effect on<br />

ERK1/2<br />

Rosiglitazone Rosiglitazone<br />

Cell growth and<br />

vascular remodeling<br />

independent of ERK1/2<br />

Conduit<br />

Artery<br />

Ang II/AT 1<br />

MEK<br />

ERK1/2<br />

Transcription<br />

factors<br />

Cell growth and<br />

vascular remodeling<br />

FIGURE 12.2 Angiotensin (Ang) II effects on extracellular-regulated kinase (ERK) 1 and<br />

2 activation in resistance and conduit arteries.<br />

signaling in aortic VSMCs. These data collectively demonstrate the countervailing<br />

effects of PPAR activators on MAPK signaling pathways implicated in cell growth<br />

and migration induced by vasoactive agents such as Ang II.<br />

p38 SIGNALING IN VASCULAR REMODELING<br />

p38 MAPK comprises six isoforms (p38α1 and α2, p38β1 and β2, p38δ, and<br />

p38γ): the α and β isoforms are ubiquitous; the δ isoform is expressed mainly in<br />

kidneys, lungs, and pancreas; and the γ isoform is highly expressed in skeletal<br />

muscle. 39 p38 MAPK can be activated by a variety of stimuli including cytokines,<br />

growth factors, osmotic shock, and different stressors. p38 MAPK has been<br />

implicated in several pathophysiological conditions such as atherosclerosis,<br />

hypertensive vascular remodeling, and ischemic cardiomyopathy. 26<br />

Ang II can stimulate vascular p38 MAPK, affecting inflammatory responses,<br />

apoptosis, cell growth inhibition, and contractility. Inhibition of p38 MAPK<br />

reduced these effects in aortic but not mesenteric VSMCs. 40–42 p38 MAPK crosstalks<br />

with other members of the MAPK family (including ERK1/2, MEK3, and<br />

MEK6; see Figure 12.1) and other signaling molecules (e.g., heat shock protein<br />

27), which is important in cytoskeletal reorganization implicated in cell migration<br />

and vascular remodeling. 26,39 p38 MAPK is also regulated by PPAR-α and one<br />

of its co-activators, PGC-1, suggesting a regulatory role for p38 MAPK in PPARα-mediated<br />

lipid metabolism. 27 However, p38 MAPK does not seem to participate<br />

in PPAR-γ regulation, as inhibition of p38 MAPK failed to modulate PPAR-γinduced<br />

AT 1 gene transcription. 43<br />

This further demonstrates differential effects of different MAPKs on the RAAS.<br />

Since p38 MAPK may also induce VSMC apoptosis, PPAR-α activators such as


Receptors in Vascular Health in Hypertension and Diabetes 201<br />

ω-3 fatty acids (e.g., docosahexaenoic acid) may stimulate VSMC apoptosis via a<br />

p38/PPAR-α-dependent manner. 44 As with ERK1/2, endogenous (retinoic acid) and<br />

synthetic PPAR ligands (various TZDs such as ciglitazone, troglitazone, and<br />

rosiglitazone) modulate p38 MAPK activity. 38,45,46 However, both ciglitazone and<br />

15d-PGJ2 are potent stimulators of p38 MAPK activity in rat astrocytes but not in<br />

mesangial cells. 45,47 These data demonstrate that PPARs have the ability to directly<br />

and indirectly regulate MAPKs in a tissue-dependent manner.<br />

ROLE OF JNK SIGNALING IN VASCULAR REMODELING<br />

The third member of the MAPK family that we will address is c-Jun N-terminal<br />

kinase (JNK) which comprises JNK1, JNK2, and JNK3. JNKs can be stimulated<br />

by several cytokines, growth factors, stressors, and vasoactive agents (Ang II and<br />

endothelin-1). 26 Contrary to the pro-stimulatory effects of Ang II on cell growth<br />

via ERK1/2, JNK activation induced apoptosis and cell cycle inhibition.<br />

Once activated, JNKs activate transcription factors such as c-Jun, c-Myc,<br />

PPAR-α, and PPAR-γ. 27,48 Ang II-induced JNK activation is dependent upon other<br />

factors such Ca 2+ mobilization and PKC-ξ activation, but is independent of c-Src<br />

activation. 49 JNK also seems to be activated in a cell-specific manner. In cardiac<br />

fibroblasts, Ang II stimulated JNK via PKC-independent proline-rich tyrosine<br />

kinase 2/Rac1, whereas in VSMCs this occurs through PAK, PKC, and Ca 2+ . 50<br />

Contrary to the other MAPK proteins, the cellular effects of PPARs on the<br />

JNK pathway remain to be fully elucidated in VSMCs. PPAR-γ activation may<br />

induce JNK signaling involved in apoptosis. Taken together, all MAPK pathways<br />

are involved in molecular mechanisms that ultimately lead to development and<br />

maintenance of cardiovascular dysfunction such as the vascular remodeling<br />

observed in hypertension and diabetes.<br />

PI3K SIGNALING PATHWAY <strong>AND</strong> VASCULAR REMODELING<br />

The PI3K pathway is implicated in cardiovascular events such as cardiac and<br />

vascular remodeling, heart failure, and endothelial dysfunction. 51,52 The PI3K<br />

family is composed of 3′-OH inositol phosphorylating enzymes and includes<br />

members of three classes divided according to their molecular structures, activation<br />

mechanisms, and substrate specificities. Class I (A and B) is composed of<br />

heterodimeric regulatory and catalytic subunits while class II proteins are composed<br />

of single catalytic subunits (PI3K-C2α, PI3K-C2β, or PI3K-C2γ). 53 Class<br />

IA has three catalytic (p110α, β, and δ) and three regulatory subunits (p85α,<br />

p85β, and p55γ) expressed mostly in the heart and blood vessels. However, class<br />

IB contains only one protein kinase composed of a p110γ catalytic subunit and<br />

a p101 regulatory subunit, expressed predominantly in cardiomyocytes, fibroblasts,<br />

endothelial cells, and VSMCs. 53<br />

The importance of PI3K in Ang II-induced cardiac hypertrophy and VSMC<br />

proliferation has been established with the use of specific reversible (LY294002)


202 Oxidative Stress and Inflammatory Mechanisms<br />

and irreversible (wortmannin) inhibitors. 51 PI3K regulates cell growth and survival,<br />

cytoskeletal reorganization, ROS production, and glucose transport among<br />

other functions and plays a role in the progression of cardiac hypertrophy and<br />

dysfunction and in vascular remodeling. 52–54 PI3K is regulated by other kinases<br />

such as Rho kinase, which may contribute to endothelial dysfunction. 55 PI3K<br />

participates in endothelial nitric oxide synthase (eNOS) activation and thus in<br />

NO generation and vasodilation. 56<br />

Ang II-stimulated ERK1/2 activity was blunted by the PI3K inhibitor LY-<br />

294009 in mesenteric artery VSMCs from hypertensive rats but not normotensive<br />

rats, which may imply cross-talk between PI3K and ERK1/2 pathways. 29 PPAR-γ<br />

regulates PI3K activity. Rosiglitazone induced PI3K/p85α expression and activity<br />

in adipocytes and in conduit and resistance arteries, but not in skeletal muscle.<br />

14,57,58 These data suggest that altered PI3K activity and expression may<br />

occur at the initiation and during the maintenance phase of hypertensive vascular<br />

disease.<br />

ROLE OF Akt/PKB SIGNALING PATHWAY IN VASCULAR<br />

REMODELING<br />

A key protein within the PI3K signaling pathway is Akt, also known as protein<br />

kinase B (PKB). In mammals, three Akt/PKB isoforms (Akt1/PKBα, Akt2/PKBβ,<br />

and Akt3/PKBγ) are highly expressed in cardiac, vascular, and endothelial cells. 53<br />

They are regulated by PI3K, SH2-containing inositol phosphatase (SHIP2) and<br />

phosphoinositide-dependent kinase (PDK)-1 and -2. 59 Once activated, Akt/PKB<br />

modulates cell survival, growth, migration, and glucose metabolism. 60–63 Synthetic<br />

and endogenous PPAR ligands inhibit Akt/PKB in blood vessels and other<br />

tissues, possibly through the inhibition of eukaryotic initiation factor-4E-binding<br />

protein (4E-BP)-1, a protein implicated in cellular growth. 14,64<br />

ROS <strong>AND</strong> VASCULAR REMODELING<br />

Reactive oxygen species (ROS) play important roles in physiological and pathophysiological<br />

conditions related to cell growth, differentiation, migration and<br />

signaling, extracellular matrix production and degradation, and inflammation in<br />

the cardiovascular system. ROS include superoxide anion (•O 2 – ), hydrogen peroxide<br />

(H 2O 2), hydroxyl radical (•OH – ), and peroxynitrite (ONOO – ) as well as<br />

other free radicals such as NO. 22,65,66<br />

Ang II is a critical regulator of ROS generation in many tissues including<br />

VSMCs and cardiac, mesangial, and endothelial cells. 22 NAD(P)H oxidase is the<br />

most important vascular source of ROS. It is composed of five subunits: p22 phox ,<br />

gp91 phox (membrane-bound), p47 phox , p67 phox , and a small G protein Rac1 or Rac2<br />

(cytosolic). In addition to NAD(P)H oxidase, other sources of ROS include<br />

xanthine oxidase, uncoupled eNOS, and the mitochondrial respiratory chain. 22<br />

ROS generation in response to pressor doses of Ang II plays a major role in the


Receptors in Vascular Health in Hypertension and Diabetes 203<br />

deleterious effects of this peptide. These effects are BP-independent effects of<br />

Ang II since similar BP elevation achieved by norepinephrine infusion did not<br />

increase vascular •O 2 – production. 67<br />

The pro-growth and hypertensive effects of ROS have also been shown to be<br />

mediated by ONOO – production via NO scavenging. 68 Many inhibitors of ROS<br />

production have been used to elucidate the detrimental role of ROS in hypertension<br />

and diabetes: protein nitration, lipid oxidation, and DNA degradation. The<br />

inhibitors include superoxide dismutase (SOD) mimetics (Tempol), •O 2 – scavengers<br />

(Tiron), specific inhibitors of NAD(P)H oxidase (apocynin, gp91ds-tat), SOD<br />

inhibitors (PEG-SOD), xanthine oxidase inhibitors (allopurinol) or mitochondrial<br />

chain inhibitors (thenotrifluoroacetone, carbonyl cyanide-m-chlorophenylhydrazone<br />

and rotenone). 69–71 Additionally, ROS interact with second messengers (e.g.,<br />

intracellular Ca 2+ ), several kinases such as MAPK (ERK, p38, and JNK) and<br />

Akt/PKB, and transcription factors (NFκB and AP-1), all leading to the development<br />

of vascular inflammation and remodeling. 22,72,73<br />

Several studies have demonstrated beneficial vascular inhibitory effects of<br />

PPAR activators on ROS production. PPAR-α activators (docosahexaenoic acid,<br />

DHA, and fenofibrate) and PPAR-γ activators (rosiglitazone or pioglitazone)<br />

reduced ROS production in rats infused with Ang II and in endothelin-1-dependent<br />

hypertensive models such as DOCA-salt hypertensive rats. 16,74<br />

VASCULAR INFLAMMATION <strong>AND</strong> REMODELING<br />

One of the major effects of activation of the signaling pathways described is the<br />

induction of vascular inflammation found in cardiovascular diseases such as<br />

hypertension and in diabetes. 75 Of the various systems involved, one of the major<br />

mechanisms is the pro-inflammatory effect of Ang II mediated via the AT 1 receptor.<br />

75,76 These effects are mediated by increased expression of adhesion molecules<br />

[intercellular adhesion molecule (ICAM)-1, platelet endothelial cell adhesion<br />

molecule (PECAM), vascular cell adhesion molecule-1 (VCAM-1), and selectins]<br />

and transcription factors (NFκB, AP-1) on monocytes, endothelial cells, and<br />

VSMCs. 13,77,78<br />

In addition to increased adhesion molecules, Ang II stimulates monocyte<br />

chemotactic protein (MCP)-1 synthesis in monocytes/macrophages, endothelial<br />

cells, and VSMCs, causing accumulation of inflammatory cells and molecules in<br />

the vasculature that ultimately contributes to the development of endothelial<br />

dysfunction and vascular remodeling. 79 PPAR activators exert potent anti-inflammatory<br />

effects. For instance, PPAR-α activation with fenofibrate and gemfibrozil<br />

inhibited cytokine production (TNFα, interferon-γ, interleukin (IL)-6, IL-2, and<br />

IL-1β) and adhesion molecule synthesis, whereas eNOS and COX-2 expressions<br />

were increased in VSMCs. 80–82<br />

These effects were associated with inhibition of expression and activity of<br />

transcription factors NFκB and AP-1. 83,84 A PPAR-α activator prevented hypertension<br />

and vascular remodeling by a reduction in NAD(P)H oxidase activity,<br />

VCAM-1 and ICAM-1 expression, in Ang II infused rats. 74 PPAR-γ also has


204 Oxidative Stress and Inflammatory Mechanisms<br />

potent anti-inflammatory properties in addition to its insulin-sensitizing effects.<br />

Among beneficial effects reported are reductions in plasminogen activator inhibitor<br />

(PAI)-1 in microalbuminuria and in matrix metalloproteinase (MMP)-9 along<br />

with increased NO generation. 13,47,85 The anti-inflammatory vascular effects of<br />

PPAR-γ occur through inhibition of NFκB and AP-1, as well as down-regulation<br />

of NAD(P)H oxidase subunits. 13,86<br />

DUAL PPAR ACTIVATORS<br />

Because of the beneficial cardiovascular effects of PPAR-α and PPAR-γ activators<br />

such as improvement of lipid metabolism, insulin sensitization, glucose metabolism,<br />

vascular remodeling and inflammation, combined activation of these<br />

nuclear receptors has been attempted. To date, at least eight dual PPARα/γ<br />

activators are being evaluated in studies ranging from preclinical to phase III.<br />

The activators have been shown to decrease circulating triglyceride concentrations<br />

in humans and transgenic human Apo A-I mice. 87,88<br />

Dual PPAR-α/γ activator effects on glucose metabolism and insulin sensitization<br />

are similar to those of PPAR-γ activators administered alone. 87,88 Two dual<br />

PPAR-α/γ activators, ragaglitazar and muraglitazar, demonstrated beneficial<br />

effects on vascular complications in hypertension and on metabolic abnormalities<br />

in diabetes, respectively. 89,90 However, muraglitazar administration in type 2 diabetic<br />

patients was associated with more death and major adverse cardiovascular<br />

events including myocardial infarction, stroke, and transient ischemic attacks,<br />

than use of a PPAR-γ activator. 91<br />

This result dampened interest in these dual PPAR activators, which may not<br />

really be superior to agents with single receptor activating capability. Because of<br />

these reported deleterious effects, we and others have used a different approach<br />

to stimulate both PPAR-α and PPAR-γ receptors. We administered concomitantly<br />

sub-therapeutic doses of PPAR-α and PPAR-γ activators that had beneficial effects<br />

when administered in full doses. At these lower doses, a combination of PPARα<br />

and PPAR-γ administration had beneficial effects in a rodent model of Ang IIinduced<br />

hypertension on vascular function, including improvement of endothelial<br />

function, decreased ROS generation, and vascular inflammation. Seber et al. also<br />

found that concomitant administration of rosiglitazone and fenofibrate improved<br />

the atherogenic dyslipidemic profiles of type II diabetic patients with poor metabolic<br />

control. 92<br />

CONCLUSION<br />

Hypertension is highly prevalent and one of the major causes of burden of disease,<br />

particularly cardiovascular morbidity and mortality. Diabetes is increasing worldwide,<br />

and often is associated with hypertension; it puts hypertensive subjects at<br />

the highest cardiovascular risk. Both conditions are associated with vascular


Receptors in Vascular Health in Hypertension and Diabetes 205<br />

injury that serves as the major mechanism for cardiovascular events that lead to<br />

myocardial infarction, stroke, amputations, and renal failure.<br />

We have summarized here some of the pathways that participate in growth,<br />

inflammation, and oxidative stress that lead to atherosclerosis and vascular remodeling<br />

in hypertension and diabetes. PPAR-α and PPAR-γ appear to act as countervailing<br />

influences on one of the major activators of the pathways that trigger<br />

vascular disease, the RAAS. Preclinical and clinical data suggest that PPAR-α<br />

and PPAR-γ activators may exert important vascular protective effects. Although<br />

initial experience with agents endowed with combined PPAR-α and PPAR-γ<br />

stimulatory effects has been disappointing, mechanistic evidence suggests that<br />

agents with these properties that are able to affect the deleterious intracellular<br />

signaling pathways described in this chapter may be developed eventually and<br />

successfully contribute to reducing the burden of disease generated by hypertension<br />

and diabetes.<br />

REFERENCES<br />

1. Haffner SM, Lehto S, Ronnemaa T, Pyorala K, and Laakso M. Mortality from<br />

coronary heart disease in subjects with type 2 diabetes and in nondiabetic subjects<br />

with and without prior myocardial infarction. New Engl J Med 1998; 339: 229.<br />

2. Schiffrin EL. Reactivity of small blood vessels in hypertension: relation with<br />

structural changes: state of the art lecture. Hypertension 1992; 19, Suppl:II1.<br />

3. Schiffrin EL. Remodeling of resistance arteries in essential hypertension and<br />

effects of antihypertensive treatment. Am J Hypertens 2004; 17: 1192.<br />

4. Heagerty AM, Aalkjaer C, Bund SJ, Korsgaard N, and Mulvany MJ. Small artery<br />

structure in hypertension: dual processes of remodeling and growth. Hypertension<br />

1993; 21: 391.<br />

5. Intengan HD, Deng LY, Li JS, and Schiffrin EL. Mechanics and composition of<br />

human subcutaneous resistance arteries in essential hypertension. Hypertension<br />

1999; 33, Pt 2: 569.<br />

6. Deng LY, Li JS, and Schiffrin EL. Endothelium-dependent relaxation of small<br />

arteries from essential hypertensive patients: mechanisms and comparison with<br />

normotensive subjects and with responses of vessels from spontaneously hypertensive<br />

rats. Clin Sci 1995; 88: 611.<br />

7. Rizzoni D et al. Structural alterations in subcutaneous small arteries of normotensive<br />

and hypertensive patients with non-insulin-dependent diabetes mellitus.<br />

Circulation 2001; 103: 1238.<br />

8. Endemann DH et al. Persistent remodeling of resistance arteries in type 2 diabetic<br />

patients on antihypertensive treatment. Hypertension 2004; 43: 399.<br />

9. Rizzoni D et al. Endothelial dysfunction in small resistance arteries of patients<br />

with non-insulin-dependent diabetes mellitus. J Hypertens 2001; 19: 913.<br />

10. United Kingdom Prospective Diabetes Study Group. Tight blood pressure control<br />

and risk of macrovascular and microvascular complications in type 2 diabetes. Br<br />

Med J 1998; 317: 703.<br />

11. Dzau VJ. Theodore Cooper Lecture: tissue angiotensin and pathobiology of vascular<br />

disease: a unifying hypothesis. Hypertension 2001; 37: 1047.


206 Oxidative Stress and Inflammatory Mechanisms<br />

12. Touyz RM and Schiffrin EL. Signal transduction mechanisms mediating the physiological<br />

and pathophysiological actions of angiotensin II in vascular smooth<br />

muscle cells. Pharmacol Rev 2000; 52: 639.<br />

13. Diep QN et al. Structure, endothelial function, cell growth, and inflammation in<br />

blood vessels of angiotensin II-infused rats: role of peroxisome proliferator-activated<br />

receptor-γ. Circulation 2002; 105: 2296.<br />

14. Benkirane K, Viel EC, Amiri F, and Schiffrin EL. Peroxisome proliferator-activated<br />

receptor γ regulates angiotensin II-stimulated phosphatidylinositol 3-kinase<br />

and mitogen-activated protein kinase in blood vessels in vivo. Hypertension 2006;<br />

47: 102.<br />

15. Diep QN, Amiri F, Benkirane K, Paradis P, and Schiffrin EL. Long-term effects<br />

of the PPARgamma activator pioglitazone on cardiac inflammation in stroke-prone<br />

spontaneously hypertensive rats. Can J Physiol Pharmacol 2004; 82: 976.<br />

16. Iglarz M, Touyz RM, Amiri F, Lavoie MF, Diep QN, and Schiffrin EL. Effect of<br />

peroxisome proliferator-activated receptor-α and -γ activators on vascular remodeling<br />

in endothelin-dependent hypertension. Arterioscler Thromb Vasc Biol 2003;<br />

23: 45.<br />

17. Sidhu JS, Kaposzta Z, Markus HS, and Kaski JC. Effect of rosiglitazone on<br />

common carotid intima-media thickness progression in coronary artery disease<br />

patients without diabetes mellitus. Arterioscler Thromb Vasc Biol 2004; 24: 930.<br />

18. Wang TD, Chen WJ, Lin JW, Chen MF, and Lee YT. Effects of rosiglitazone on<br />

endothelial function, C-reactive protein, and components of the metabolic syndrome<br />

in nondiabetic patients with the metabolic syndrome. Am J Cardiol 2004;<br />

93: 362.<br />

19. Haffner SM, Greenberg AS, Weston WM, Chen H, Williams K, and Freed MI.<br />

Effect of rosiglitazone treatment on nontraditional markers of cardiovascular disease<br />

in patients with type 2 diabetes mellitus. Circulation 2002; 106: 679.<br />

20. Schiffrin EL. Peroxisome proliferator-activated receptors and cardiovascular<br />

remodeling. Am J Physiol Heart Circ Physiol 2005; 288: H1037.<br />

21. Berger J and Moller DE. The mechanisms of action of PPARs. Annu Rev Med<br />

2002; 53: 409.<br />

22. Touyz RM and Schiffrin EL. Reactive oxygen species in vascular biology: implications<br />

in hypertension. Histochem Cell Biol 2004; 122: 339.<br />

23. Robinson MJ and Cobb MH. Mitogen-activated protein kinase pathways. Curr<br />

Opin Cell Biol 1997; 9: 180.<br />

24. Touyz RM, He G, Deng LY, and Schiffrin EL. Role of extracellular signalregulated<br />

kinases in angiotensin II-stimulated contraction of smooth muscle cells<br />

from human resistance arteries. Circulation 1999; 99: 392.<br />

25. Touyz RM, He G, El Mabrouk M, Diep Q, Mardigyan V, and Schiffrin EL.<br />

Differential activation of extracellular signal-regulated protein kinase 1/2 and p38<br />

mitogen-activated protein kinase by AT1 receptors in vascular smooth muscle cells<br />

from Wistar-Kyoto rats and spontaneously hypertensive rats. J Hypertens 2001;<br />

19: 553.<br />

26. Pearson G et al. Mitogen-activated protein (MAP) kinase pathways: regulation<br />

and physiological functions. Endocr Rev 2001; 22: 153.<br />

27. Yang SH, Sharrocks AD, and Whitmarsh AJ. Transcriptional regulation by the<br />

MAP kinase signaling cascades. Gene 2003; 320: 3.<br />

28. Fiebeler A and Haller H. Participation of the mineralocorticoid receptor in cardiac<br />

and vascular remodeling. Nephron Physiol 2003; 94: 47.


Receptors in Vascular Health in Hypertension and Diabetes 207<br />

29. El Mabrouk M, Touyz RM, and Schiffrin EL. Differential ANG II-induced growth<br />

activation pathways in mesenteric artery smooth muscle cells from SHR. Am J<br />

Physiol Heart Circ Physiol 2001; 281: H30.<br />

30. Kohlstedt K and Busse R, Fleming I. Signaling via the angiotensin-converting<br />

enzyme enhances the expression of cyclooxygenase-2 in endothelial cells. Hypertension<br />

2005; 45: 126.<br />

31. Kintscher U et al. Angiotensin II induces migration and Pyk2/paxillin phosphorylation<br />

of human monocytes. Hypertension 2001; 37: 587.<br />

32. Touyz RM. Recent advances in intracellular signalling in hypertension. Curr Opin<br />

Nephrol Hypertens 2003; 12: 165.<br />

33. Kim J et al. Mitogen-activated protein kinase contributes to elevated basal tone<br />

in aortic smooth muscle from hypertensive rats. Eur J Pharmacol 2005; 514: 209.<br />

34. Rice KM, Kinnard RS, Harris R, Wright GL, and Blough ER. Effects of aging<br />

on pressure-induced MAPK activation in the rat aorta. Pflugers Arch 2005; 450:<br />

192.<br />

35. Zhang B et al. Insulin- and mitogen-activated protein kinase-mediated phosphorylation<br />

and activation of peroxisome proliferator-activated receptor gamma. J<br />

Biol Chem 1996; 271: 31771.<br />

36. Verlohren S et al. Visceral periadventitial adipose tissue regulates arterial tone of<br />

mesenteric arteries. Hypertension 2004; 44: 271.<br />

37. Rao GN, Alexander RW, and Runge MS. Linoleic acid and its metabolites, hydroperoxyoctadecadienoic<br />

acids, stimulate c-Fos, c-Jun, and c-Myc mRNA expression,<br />

mitogen-activated protein kinase activation, and growth in rat aortic smooth<br />

muscle cells. J Clin Invest 1995; 96: 842.<br />

38. Teruel T, Hernandez R, Benito M, and Lorenzo M. Rosiglitazone and retinoic<br />

acid induce uncoupling protein-1 (UCP-1) in a p38 mitogen-activated protein<br />

kinase-dependent manner in fetal primary brown adipocytes. J Biol Chem 2003;<br />

278: 263.<br />

39. McMullen ME, Bryant PW, Glembotski CC, Vincent PA, and Pumiglia KM.<br />

Activation of p38 has opposing effects on the proliferation and migration of<br />

endothelial cells. J Biol Chem 2005; 280: 20995.<br />

40. Ushio-Fukai M, Alexander RW, Akers M, and Griendling KK. p38 Mitogenactivated<br />

protein kinase is a critical component of the redox-sensitive signaling<br />

pathways activated by angiotensin II: role in vascular smooth muscle cell hypertrophy.<br />

J Biol Chem 1998; 273: 15022.<br />

41. Meloche S, Landry J, Huot J, Houle F, Marceau F, and Giasson E. p38 MAP<br />

kinase pathway regulates angiotensin II-induced contraction of rat vascular smooth<br />

muscle. Am J Physiol Heart Circ Physiol 2000; 279: H741.<br />

42. Touyz RM, He G, El Mabrouk M, and Schiffrin EL. p38 Map kinase regulates<br />

vascular smooth muscle cell collagen synthesis by angiotensin II in SHR but not<br />

in WKY. Hypertension 2001; 37: 574.<br />

43. Sugawara A, Takeuchi K, Uruno A, Kudo M, Sato K, and Ito S. Effects of mitogenactivated<br />

protein kinase pathway and co-activator CREP-binding protein on peroxisome<br />

proliferator-activated receptor-gamma-mediated transcription suppression<br />

of angiotensin II type 1 receptor gene. Hypertens Res 2003; 26: 623.<br />

44. Diep QN, Touyz RM, and Schiffrin EL. Docosahexaenoic acid, a peroxisome<br />

proliferator-activated receptor-alpha ligand, induces apoptosis in vascular smooth<br />

muscle cells by stimulation of p38 mitogen-activated protein kinase. Hypertension<br />

2000; 36: 851.


208 Oxidative Stress and Inflammatory Mechanisms<br />

45. Lennon AM, Ramauge M, Dessouroux A, and Pierre M. MAP kinase cascades<br />

are activated in astrocytes and preadipocytes by 15-deoxy-delta(12-14)-prostaglandin<br />

J(2) and the thiazolidinedione ciglitazone through peroxisome proliferator<br />

activator receptor gamma-independent mechanisms involving reactive oxygenated<br />

species. J Biol Chem 2002; 277: 29681.<br />

46. Gardner OS, Shiau CW, Chen CS, and Graves LM. Peroxisome proliferatoractivated<br />

receptor gamma-independent activation of p38 MAPK by thiazolidinediones<br />

involves calcium/calmodulin-dependent protein kinase II and protein kinase<br />

R: correlation with endoplasmic reticulum stress. J Biol Chem 2005; 280: 10109.<br />

47. Hsueh WA and Law RE. PPAR- and atherosclerosis: effects on cell growth and<br />

movement. Arterioscler Thromb Vasc Biol 2001; 21: 1891.<br />

48. Diradourian C, Girard J, and Pegorier JP. Phosphorylation of PPARs: from molecular<br />

characterization to physiological relevance. Biochimie 2005; 87: 33.<br />

49. Liao DF, Monia B, Dean N, and Berk BC. Protein kinase C-zeta mediates angiotensin<br />

II activation of ERK1/2 in vascular smooth muscle cells. J Biol Chem 1997;<br />

272: 6146.<br />

50. Murasawa S et al. Angiotensin II initiates tyrosine kinase Pyk2-dependent signalings<br />

leading to activation of Rac1-mediated c-Jun NH2-terminal kinase. J Biol<br />

Chem 2000; 275: 26856.<br />

51. Rocic P, Govindarajan G, Sabri A, and Lucchesi PA. A role for PYK2 in regulation<br />

of ERK1/2 MAP kinases and PI 3-kinase by ANG II in vascular smooth muscle.<br />

Am J Physiol Cell Physiol 2001; 280: C90.<br />

52. Anderson KE and Jackson SP. Class I phosphoinositide 3-kinases. Int J Biochem<br />

Cell Biol 2003; 35: 1028.<br />

53. Oudit GY et al. The role of phosphoinositide-3 kinase and PTEN in cardiovascular<br />

physiology and disease. J Mol Cell Cardiol 2004; 37: 449.<br />

54. Hafizi S, Wang X, Chester AH, Yacoub MH, and Proud CG. ANG II activates<br />

effectors of mTOR via PI3-K signaling in human coronary smooth muscle cells.<br />

Am J Physiol Heart Circ Physiol 2004; 287: H1232.<br />

55. Budzyn K, Marley PD, and Sobey CG. Opposing roles of endothelial and smooth<br />

muscle phosphatidylinositol 3-kinase in vasoconstriction: effects of rho-kinase<br />

and hypertension. J Pharmacol Exp Ther 2005; 313: 1248.<br />

56. Andreozzi F, Laratta E, Sciacqua A, Perticone F, and Sesti G. Angiotensin II<br />

impairs the insulin signaling pathway promoting production of nitric oxide by<br />

inducing phosphorylation of insulin receptor substrate-1 on Ser312 and Ser616<br />

in human umbilical vein endothelial cells. Circ Res 2004; 94: 1211.<br />

57. Rieusset J et al. The expression of the p85alpha subunit of phosphatidylinositol<br />

3-kinase is induced by activation of the peroxisome proliferator-activated receptor<br />

gamma in human adipocytes. Diabetologia 2001; 44: 544.<br />

58. Rieusset J, Roques M, Bouzakri K, Chevillotte E, and Vidal H. Regulation of p85alpha<br />

phosphatidylinositol-3-kinase expression by peroxisome proliferator-activated<br />

receptors (PPARs) in human muscle cells. FEBS Lett 2001; 502: 98.<br />

59. Dong LQ and Liu F. PDK2: the missing piece in the receptor tyrosine kinase<br />

signaling pathway puzzle. Am J Physiol Endocrinol Metab 2005; 289: E187.<br />

60. Shiojima I and Walsh K. Role of Akt signaling in vascular homeostasis and<br />

angiogenesis. Circ Res 2002; 90: 1243.<br />

61. Shioi T et al. Akt/protein kinase B promotes organ growth in transgenic mice.<br />

Mol Cell Biol 2002; 22: 2799.


Receptors in Vascular Health in Hypertension and Diabetes 209<br />

62. Li F, Zhang C, Schaefer S, Estes A, and Malik KU. ANG II-induced neointimal<br />

growth is mediated via c. Am J Physiol Heart Circ Physiol 2005; 289: H2592.<br />

63. Ohashi H et al. Phosphatidylinositol 3-kinase/Akt regulates angiotensin II-induced<br />

inhibition of apoptosis in microvascular endothelial cells by governing survivin<br />

expression and suppression of caspase-3 activity. Circ Res 2004; 94: 785.<br />

64. Benkirane K, Amiri F, Diep QN, El Mabrouk M, and Schiffrin EL. PPAR-γ inhibits<br />

ANG II-induced cell growth via SHIP2 and 4E-BP1. Am J Physiol Heart Circ<br />

Physiol 2006; 290: H390.<br />

65. Griendling KK, Sorescu D, Lassegue B, and Ushio-Fukai M. Modulation of<br />

protein kinase activity and gene expression by reactive oxygen species and their<br />

role in vascular physiology and pathophysiology. Arterioscler Thromb Vasc Biol<br />

2000; 20: 2175.<br />

66. Paravicini TM, and Touyz RM. Redox signaling in hypertension. Cardiovasc Res<br />

2006; 71: 247.<br />

67. Rajagopalan S et al. Angiotensin II-mediated hypertension in the rat increases<br />

vascular superoxide production via membrane NADH/NADPH oxidase activation:<br />

contribution to alterations of vasomotor tone. J Clin Invest 1996; 97: 1916.<br />

68. Muijsers RB, Folkerts G, Henricks PA, Sadeghi-Hashjin G, and Nijkamp FP.<br />

Peroxynitrite: a two-faced metabolite of nitric oxide. Life Sci 1997; 60: 1833.<br />

69. Park JB, Touyz RM, Chen X, and Schiffrin EL. Chronic treatment with a superoxide<br />

dismutase mimetic prevents vascular remodeling and progression of hypertension<br />

in salt-loaded stroke-prone spontaneously hypertensive rats. Am J<br />

Hypertens 2002; 15: 78.<br />

70. Rey FE, Cifuentes ME, Kiarash A, Quinn MT, and Pagano PJ. Novel competitive<br />

inhibitor of NAD(P)H oxidase assembly attenuates vascular O 2 – and systolic blood<br />

pressure in mice. Circ Res 2001; 89: 408.<br />

71. Touyz RM, Yao G, Viel E, Amiri F., and Schiffrin EL. Angiotensin II and endothelin-1<br />

regulate MAP kinases through different redox-dependent mechanisms in<br />

human vascular smooth muscle cells. J Hypertens 2004; 22: 1141.<br />

72. Torrecillas G et al. The role of hydrogen peroxide in the contractile response to<br />

angiotensin II. Mol Pharmacol 2001; 59: 104.<br />

73. Browatzki M et al. Angiotensin II stimulates matrix metalloproteinase secretion<br />

in human vascular smooth muscle cells via nuclear factor-kappaB and activator<br />

protein 1 in a redox-sensitive manner. J Vasc Res 2005; 42: 415.<br />

74. Diep QN, Amiri F, Touyz RM, Cohn JS, Endemann D, Neves MF, and Schiffrin<br />

EL. PPAR-α activator effects on Ang II-induced vascular oxidative stress and<br />

inflammation. Hypertension 2002; 40: 866.<br />

75. Savoia C and Schiffrin EL. Inflammation in hypertension. Curr Opin Nephrol<br />

Hypertens 2006; 15(2):152-158.<br />

76. Victorino GP, Newton CR, and Curran B. Effect of angiotensin II on microvascular<br />

permeability. J Surg Res 2002; 104: 77.<br />

77. Grafe M et al. Angiotensin II-induced leukocyte adhesion on human coronary<br />

endothelial cells is mediated by E-selectin. Circ Res 1997; 81: 804.<br />

78. Pueyo ME et al. Angiotensin II stimulates endothelial vascular cell adhesion<br />

molecule-1 via nuclear factor-κB activation induced by intracellular oxidative<br />

stress. Arterioscler Thromb Vasc Biol 2000; 20: 645.<br />

79. Cheng ZJ, Vapaatalo H, and Mervaala E. Angiotensin II and vascular inflammation.<br />

Med Sci Monit 2005; 11: RA194.


210 Oxidative Stress and Inflammatory Mechanisms<br />

80. Moraes LA, Piqueras L, and Bishop-Bailey D. Peroxisome proliferator-activated<br />

receptors and inflammation. Pharmacol Ther 2006; 110: 371.<br />

81. Forman BM, Chen J, and Evans RM. Hypolipidemic drugs, polyunsaturated fatty<br />

acids, and eicosanoids are ligands for peroxisome proliferator-activated receptors<br />

α and δ. Proc Natl Acad Sci USA 1997; 94: 4312.<br />

82. Hu ZW, Kerb R, Shi XY, Wei-Lavery T, and Hoffman BB. Angiotensin II increases<br />

expression of cyclooxygenase-2: implications for the function of vascular smooth<br />

muscle cells. J Pharmacol Exp Ther 2002; 303: 563.<br />

83. Delerive P et al. Peroxisome proliferator-activated receptor α negatively regulates<br />

the vascular inflammatory gene response by negative cross-talk with transcription<br />

factors NF-κB and AP-1. J Biol Chem 1999; 274: 32048.<br />

84. Staels B et al. Activation of human aortic smooth-muscle cells is inhibited by<br />

PPAR- but not by PPAR- activators. Nature 1998; 393: 790.<br />

85. Martens FM et al. Metabolic and additional vascular effects of thiazolidinediones.<br />

Drugs 2002; 62: 1463.<br />

86. Hwang J et al. Peroxisome proliferator-activated receptor-γ ligands regulate endothelial<br />

membrane superoxide production. Am J Physiol Cell Physiol 2005; 288:<br />

C899.<br />

87. Lohray BB et al. (-)3-[4-[2-(Phenoxazin-10-yl)ethoxy]phenyl]-2-ethoxypropanoic<br />

acid [(-)DRF 2725]: a dual PPAR agonist with potent antihyperglycemic and lipid<br />

modulating activity. J Med Chem 2001; 44: 2675.<br />

88. Etgen GJ et al. A tailored therapy for the metabolic syndrome: the dual peroxisome<br />

proliferator-activated receptor-alpha/gamma agonist LY465608 ameliorates insulin<br />

resistance and diabetic hyperglycemia while improving cardiovascular risk<br />

factors in preclinical models. Diabetes 2002; 51: 1083.<br />

89. Mamnoor PK et al. Antihypertensive effect of ragaglitazar: a novel PPARα and<br />

γ dual activator. Pharmacol Res 2006; 54: 129.<br />

90. Harrity T et al. Muraglitazar, a novel dual (α/γ) peroxisome proliferator-activated<br />

receptor, improves diabetes and other metabolic abnormalities and perserves<br />

beta-cell function in db/db mice. Diabetes 2006; 55: 240.<br />

91. Nissen SE, Wolski K, and Topol EJ. Effect of muraglitazar on death and major<br />

adverse cardiovascular events in patients with type 2 diabetes mellitus. JAMA<br />

2005; 294: 2581.<br />

92. Seber S, Ucak S, Basat O, and Altuntas Y. The effect of dual PPAR α/γ stimulation<br />

with combination of rosiglitazone and fenofibrate on metabolic parameters in type<br />

2 diabetic patients. Diabetes Res Clin Pract 2006; 71: 52.


13<br />

CONTENTS<br />

Role of Uncoupling<br />

Protein 2 in Pancreatic β<br />

Cell Function: Secretion<br />

and Survival<br />

Jingyu Diao, Catherine B. Chan, and<br />

Michael B. Wheeler<br />

Abstract .............................................................................................................211<br />

Introduction .......................................................................................................211<br />

Regulatory Factors of UCP2 Expression in Pancreatic β Cells ......................212<br />

UCP2 Action in Insulin Secretion from Pancreatic β Cells:<br />

Role of ATP .............................................................................................216<br />

UCP2 Action in β Cell Mass and Survival:<br />

Role of ATP and ROS Production ..........................................................218<br />

Conclusion.........................................................................................................219<br />

Acknowledgments .............................................................................................219<br />

References .........................................................................................................220<br />

ABSTRACT<br />

Uncoupling protein (UCP) 2 has been considered a negative modulator of insulin<br />

secretion. In response to stress stimuli such as hyperlipidemia and inflammation,<br />

the pancreatic β cell up-regulates UCP2 expression, which results in decreased<br />

insulin secretion. Fatty acids and superoxide regulate UCP2 activity. In addition<br />

to influencing insulin secretion, UCP2 may play a role in β cell survival and<br />

proliferation.<br />

INTRODUCTION<br />

Regulation of cellular adenosine triphosphate (ATP) production and rapid adjustments<br />

in ATP levels are essential for most cells during exposure to growth stimuli<br />

211


212 Oxidative Stress and Inflammatory Mechanisms<br />

or stressful conditions such as fuel deficiency, oxidative stimuli, and inflammation.<br />

ATP is created by ATP synthase in mitochondria through coupling of the<br />

proton motive force (PMF) or electrochemical H + gradient to the oxidative phosphorylation<br />

of fuel substrates such as fatty acids or glucose. To guarantee continuous<br />

generation of ATP as the main source of energy for cellular function and<br />

viability, mitochondria operate an oxidative phosphorylation system to supply a<br />

proton gradient. In this system, PMF is generated during respiration by the<br />

passage of protons against their gradient through the electron transport chain<br />

enzyme complexes. Thus, many factors contribute directly to the regulation of<br />

PMF and thereby ATP synthesis, particularly the carrier proteins in the respiratory<br />

chain complexes in the inner membranes of mitochondria.<br />

To maintain an appropriate level of ATP production, an uncoupling process<br />

that dissipates the proton gradient and prevents the PMF from becoming excessive<br />

can be recruited. This uncoupling process also plays a role in reducing the level<br />

of reactive oxygen species (ROS), a by-product of oxidative phosphorylation<br />

produced by the complexes of the electron transport chain. 1,2 Mitochondrial<br />

uncoupling is dominated by uncoupling proteins located in the inner mitochondrial<br />

membranes. 3 Five UCP homologues in mammals sharing distinctive UCP<br />

signature sequences are thought to be involved in fatty acid anion binding and<br />

translocation. 1,4 Based on studies of UCP1, the uncoupling activity is augmented<br />

by fatty acids and inhibited by the purine nucleotide GDP. 5,6<br />

Specific uncoupling proteins play a well documented, physiological role in<br />

thermogenesis regulation. UCP1, an uncoupling protein mainly expressed in brown<br />

adipose tissue, which has little ATP synthase, induces proton leak and reduces ATP<br />

production, thereby generating heat upon cold exposure and in diet-induced thermogenesis.<br />

7,8 Activation of UCP1 stimulates respiration, fatty acid oxidation, and<br />

uncoupling activity. 9 UCP2 and UCP3 share over 50% protein sequence identity<br />

with UCP1. However, the function and regulation of these UCPs remain unclear.<br />

UCP3 is expressed in skeletal muscle and brown fat, while UCP2 is found in many<br />

tissues including heart, brain, kidney, liver, lymphocytes, pancreas, and white adipose<br />

tissue. 10 The diverse distribution of UCP2 and UCP3 led to the hypothesis<br />

that their primary physiological function involves something other than the control<br />

of thermogenesis. 11,12 Specifically, UCP2 appears to play roles in the regulation of<br />

ROS production, inflammation, and cell proliferation and death. 13,14<br />

Many recent reviews have discussed the possible physiological and pathophysiological<br />

roles of UCPs. 15–18 Thus, this chapter focuses on the role of UCP2<br />

in pancreatic β cell function including the regulation of insulin secretion and cell<br />

survival.<br />

REGULATORY FACTORS OF UCP2 EXPRESSION IN<br />

PANCREATIC β CELLS<br />

Mitochondrial function is crucial for insulin secretion. 19 For instance, blockade<br />

of the respiratory chain by mitochondrial toxins inhibits glucose-stimulated


Role of Uncoupling Protein 2 in Pancreatic β-Cell Function 213<br />

Islet Islet cells<br />

FIGURE 13.1 UCP2 protein expression in human islets. The islets were fixed with 4%<br />

paraformaldehyde, permeabilized with 0.3% Triton X-100 detergent and dispersed islet<br />

cells, and stained with goat antibody specific for UCP2 (Santa Cruz Biotechnology, Inc.).<br />

Cells were then examined using a Zeiss LSM510 confocal microscope (×40 ×1.3 magnification)<br />

and images were analyzed using LSM510 image browser software.<br />

insulin secretion. 20 Absence of specific mitochondrial proteins such as nicotinamide<br />

nucleotide transhydrogenase (Nnt) in β cells ATP production and insulin<br />

secretion. 21 Based on the association of the UCP2 gene with obesity and hyperinsulinemia,<br />

numerous reports have demonstrated the important role of UCP2 in<br />

the development of type 2 diabetes in humans. 22<br />

A common –866G/A polymorphism in the promoter of the UCP2 gene<br />

contributes to diabetes susceptibility in different human populations, perhaps due<br />

to its overall effects on pancreatic islets, the immune system, and lipid metabolism.<br />

23,24 Individuals who are heterozygous or homozygous for the –866A allele<br />

have greater UCP2 activity, which correlates with reduced glucose-induced insulin<br />

secretion and decreased oxidative stress. 25–27 Immunofluorescence staining of<br />

UCP2 revealed relatively high amounts of the UCP2 protein in isolated human<br />

islets, suggesting its importance in islet function (Figure 13.1). However, evidence<br />

linking any mutations in the UCP2 protein sequence to the pathogenesis of obesity<br />

or type 2 diabetes 28 is insufficient.<br />

In contrast, a considerable amount is known regarding the regulation of UCP2<br />

gene expression. The promoter region of UCP2 gene contains peroxisome


214 Oxidative Stress and Inflammatory Mechanisms<br />

Nutrient Deficiency<br />

SIRT1<br />

Reactive<br />

Oxygen Species<br />

?<br />

Insulin Resistance<br />

Inflammatory<br />

cytokines<br />

UCP2<br />

transcriptional/post-transcriptional regulation<br />

Nutrient Excess<br />

FIGURE 13.2 Proposed pathways of UCP2 regulation in β cells. UCP2 activity can be<br />

transcriptionally down-regulated by SIRT1 in response to nutrient deficiency and upregulated<br />

by signals from ROS, FFAs, and certain pro-inflammatory cytokines due to<br />

nutrient overload and related insulin resistance. UCP2 can also be activated directly by<br />

many factors such as ROS and FFA.<br />

proliferator response elements and a sterol regulatory element. 29–31 In pancreatic<br />

β cells, UCP2 gene expression is partially controlled by PPAR-γ coactivator PGC-<br />

1α (Peroxisome Proliferator-Activated Receptor Coactivator 1 Alpha) through<br />

sterol regulatory element binding protein isoforms (SREBPs). 32,33 Increased activity<br />

of these transcriptional activators leads to increased UCP2 expression, whereas<br />

another transcription regulator, SIRT1 (sirtuin1), represses its expression in β<br />

cells. 34<br />

SIRT1, a mammalian Sir2 orthologue, is a key regulator implicated in cellular<br />

stress response and survival through regulation of p53, NF-κB (Nuclear Factor<br />

KappaB) signaling, and FOXO (Forkhead Box O) transcription factors. 35 A moderate<br />

reduction in calorie intake (caloric restriction, CR) has been suggested to<br />

slow aging, reduce age-related chronic diseases, and extend lifespans. In simple<br />

model organisms, the sir2 gene is a strong candidate to regulate CR. 36 In mammalian<br />

cells, SIRT1 deacetylase can also be induced by CR, possibly to increase<br />

the long-term function and survival of critical cell types. 37<br />

Interestingly, SIRT1 can directly bind to the promoter of UCP2 in response<br />

to starvation in pancreatic β cells, although it acts on PPAR-γ as well. 34,38 Therefore,<br />

a decrease in SIRT1 activity in β cells would enhance UCP2 expression and<br />

reduce insulin secretion, for instance, in response to CR (Figure 13.2). As one<br />

would expect, the levels of UCP2 are low in sirt1 transgenic mice but high in<br />

sirt1 knockout mice. 34,39 It is likely, under starvation conditions, that SIRT1<br />

increases to maintain β cell survival and the cell cycle, leading to the suppression<br />

of UCP2 expression. Although food deprivation increased UCP2 mRNA in mouse<br />

pancreas, 34 cells under serum and glucose withdrawal for 15 hr increased SIRT1,<br />

possibly leading to the suppression of UCP2. 40 Conversely, under obesity or<br />

FFA<br />

FFA


Role of Uncoupling Protein 2 in Pancreatic β-Cell Function 215<br />

Untreated<br />

TNF IL-1 Palmitate C2-Ceramide<br />

FIGURE 13.3 Induction of UCP2 protein expression in pancreatic β cell MIN6. Pancreatic<br />

clonal MIN6 β cells were treated with cytokines (TNF or IL-1, 100 ng/ml), palmitate<br />

(1 mM), or ceramide (10 μM) for 48 hr, then fixed, permeabilized, and immunofluorescently<br />

stained with goat anti-UCP2 as indicated in Figure 13.1.<br />

caloric overload conditions, low levels of SIRT1 result in release of suppression<br />

on the UCP2 gene, leading to an increase of UCP2 expression. Clearly, more<br />

research is needed to determine the important link between UCPs and SIRT1 in<br />

islets.<br />

Our studies have shown that UCP2 is the predominant isoform expressed in<br />

murine pancreatic islets, with UCP1 and UCP3 present only at very low levels<br />

as demonstrated by real-time PCR (unpublished data). Interestingly, in genetically<br />

obese animal models, animals fed high fat diets, and in clonal MIN6 β cells<br />

exposed to ROS, free fatty acids (FFAs), or inflammatory cytokines, UCP2<br />

expression was up-regulated (Figure 13.3), implying that induction of UCP2 in<br />

β cells directly or indirectly correlates with insulin resistance. 41,42<br />

This is further supported by our unpublished results showing that the attenuation<br />

of insulin signaling by chronic silencing of the insulin receptor in pancreatic<br />

cells increases the expression of UCP2 protein and augments mitochondrial<br />

uncoupling activity. Together, these data suggest that β cells may adapt to obesity<br />

and insulin resistance by increasing UCP2 levels and thus uncoupling activity,<br />

consistent with the notion that UCP2 expression adapts to the metabolic state of<br />

an organism in a tissue-specific manner. 43 This adaptation with insulin resistance<br />

would result in the attenuation of insulin secretion from β cells in response to<br />

increased glucose (Figure 13.2).<br />

Because UCP2 is up-regulated in β cells in response to insulin resistant states,<br />

identifying signaling pathways regulating its expression will facilitate understanding<br />

its function at the molecular level. Along these lines, we have observed that<br />

interleukin 1(IL-1), a primary regulator of inflammatory and immune responses,<br />

increases UCP2 protein expression in mouse clonal MIN6 β cells without causing<br />

significant cytotoxic effects (Figure 13.3). In contrast, in rat clonal INS-1 β cells,<br />

IL-1β down-regulated UCP2 mRNA accompanied by decreased cell viability,


216 Oxidative Stress and Inflammatory Mechanisms<br />

suggesting that IL-1 can differently regulate UCP2 expression in β cells depending<br />

on cell signaling status. 44<br />

IL-1 binds to cell surface-specific receptors, activating specific protein kinases<br />

such as NIK (NF-κB Induced Kinase) and MAPK (Mitogen-Activated Protein<br />

Kinase), and modulates transcription factors including NF-κB, AP-1 (Adaptorrelated<br />

Protein Complex 1) and CREB (cAMP Responsive Binding Protein 1),<br />

resulting in the expression of immediate early genes central to the inflammatory<br />

response. 45 Therefore, the fact that UCP2 protein expression can be up-regulated<br />

by IL-1 suggests that UCP2 is one of the acute response genes in β cells.<br />

Consistent with this finding, IL-1β is a key inflammatory mediator causing pancreatic<br />

islet dysfunction and apoptosis through the up-regulation of inducible<br />

nitric oxide synthase and cyclooxygenase-2. 46 Although IL-1 is produced mainly<br />

by monocytes, macrophages, and other cell types, human β cells make IL-1β in<br />

response to high glucose concentrations independently of an immune-mediated<br />

process. However, the β cells also express IL-1 receptor antagonist (IL-1Ra), a<br />

naturally occurring anti-inflammatory cytokine, to protect themselves from glucotoxicity.<br />

47,48 The pathway of IL-1-mediated β cell UCP2 regulation is therefore<br />

a potential target for controlling β cell function and survival.<br />

UCP2 ACTION IN INSULIN SECRETION FROM<br />

PANCREATIC β CELLS: ROLE OF ATP<br />

The β cell is essential for glucose homeostasis due to its ability to secrete<br />

insulin in response to nutrients. 49 In general, glucose metabolism in ß cells<br />

leads to an elevated intracellular ATP:ADP ratio, which closes the ATP-sensitive<br />

K + (K ATP) channel, resulting in depolarization of the cell membranes. Closure<br />

of K ATP channels activates voltage-dependent Ca 2+ channels, leading to Ca 2+<br />

entry and insulin exocytosis. 50 Thus, given the key role ATP plays in mediating<br />

insulin secretion and synthesis, it is not surprising that UCP2 negatively regulates<br />

insulin secretion since its uncoupling activity reduces mitochondrial ATP<br />

production.<br />

Our laboratory and others have shown that cellular ATP levels in UCP2 –/–<br />

animals are elevated in association with enhanced insulin secretion, 42,51 while<br />

decreased ATP levels and impaired glucose-stimulated insulin secretion occur<br />

when UCP2 is overexpressed. 42,52,53 Furthermore, it appears that increased islet<br />

UCP2 levels occur in several rodent models of type 2 diabetes and β cell dysfunction,<br />

including hyperglycemic HFD-fed mice, ob/ob mice, fa/fa rats, mice<br />

overexpressing SREBP-1c in β cells, and rodents expressing a dominant-negative<br />

IGF-1 receptor in skeletal muscle. 42,53a–57<br />

In addition to transcriptional regulation of UCP2 expression, regulatory mechanisms<br />

that occur more rapidly are important because they allow cells to rapidly<br />

adjust ATP levels in response to various stimuli or stresses. Post-transcriptional<br />

mechanisms regulate UCP2 activity, suggesting that UCP2 is able to respond<br />

acutely to stimuli. 58 Furthermore FFAs and superoxide radicals directly activate


Role of Uncoupling Protein 2 in Pancreatic β-Cell Function 217<br />

Mitochondrial<br />

membrane potential<br />

(Safranin red (RFU))<br />

40000<br />

30000<br />

20000<br />

10000<br />

Gl-3-P 7.5mM<br />

OA 25µM<br />

OA 75µM total<br />

OA 50µM total<br />

FCCP 5.6µM<br />

0 100 200 300 400 500<br />

600<br />

Time (sec)<br />

Control<br />

UCP2<br />

UCP1<br />

Depol. and<br />

Uncoupling<br />

Hyperpol.<br />

FIGURE 13.4 FFA activates mitochondrial uncoupling. Pancreatic clonal MIN6 β cells<br />

were infected with recombinant adenovirus expressing UCP1, UCP2, and control green<br />

fluorescence protein for 48 hr, then permeabilized and subjected to measurement of<br />

mitochondrial membrane potential in a buffer containing 2.5 μM safranin. Mitochondrial<br />

membrane potential as reflected by the fluorescence was monitored by a FluoroCount plate<br />

reader at excitation/emission wavelengths of 530/590 nm. Respiratory substrate glycerol-<br />

3-phosphate (Gl-3-P, 7.5 μM) was first added to induce mitochondrial hyperpolarization.<br />

To induce OA-mediated uncoupling effects, free fatty acid oleate (OA) was added to build<br />

up OA concentration from 25 to 75 μM. To confirm the depolarization of mitochondrial<br />

membrane, carbonyl cyanide p-trifluoromethoxyphenylhydrazone (FCCP) was added to a<br />

final concentration of 5.6 μM. Vertical arrows denote time points of chemical additions.<br />

(Courtesy of Vasilij Koshkin, Wheeler Laboratory.)<br />

uncoupling. Mitochondrial superoxide activates UCP2-mediated proton leaks and<br />

influences ATP production and insulin secretion in β cells. 59–64<br />

Using safranin fluorescence to measure mitochondrial membrane potential,<br />

we demonstrated that 50 to 75 μM oleate acutely induced UCP2-mediated uncoupling<br />

within seconds in MIN6 cells, suggesting that UCP2 activity can be induced<br />

by FFAs directly or by an immediate consequence of ROS generation and lipid<br />

peroxidation of FFAs (Figure 13.4). Although the mechanism of FFA-induced<br />

UCP2 activation remains unclear, it presumably occurs through direct binding of<br />

FFA to UCP2. 1<br />

The links of UCP activators and their acute uncoupling effects on mitochondria<br />

are interesting and suggest an avenue for future investigations that may shed<br />

light on the function of UCP2. It should be noted that acute exposure of pancreatic<br />

cells to both high glucose concentrations and saturated FFAs stimulates appreciable<br />

insulin secretion, whereas chronic exposure results in desensitization and<br />

suppression of secretion. 65 Binding of FFA to its receptor GPR40 would increase<br />

intracellular cAMP levels and stimulate Ca 2+ influx through voltage-dependent<br />

Ca 2+ channels, resulting in an increase in insulin secretion. 66


218 Oxidative Stress and Inflammatory Mechanisms<br />

The effect from the FFA–GPR40 pathway may override the effect from<br />

FFA–UCP2 activity initially, considering that uncoupling has a negative impact<br />

on FFA-stimulated insulin secretion. However, it is also possible that UCP2 may<br />

initially promote insulin secretion but may ultimately attenuate it. Which of these<br />

opposing effects predominates? This is difficult to determine since UCP2 protein<br />

levels do not necessarily reflect levels of activity. Specifically, UCP2 appears to<br />

require activation by fatty acids, and studies demonstrating a quantitative correlation<br />

between UCP2 activity and insulin secretion in vivo have yet to be performed.<br />

Nevertheless, it is conceivable that in metabolic states where fuel is<br />

plentiful, increased UCP2 expression and activity by factors such as pro-inflammatory<br />

cytokines and high circulating FFA levels are highly correlated to the<br />

desensitization of β cells to release insulin.<br />

UCP2 ACTION IN ββ CELL MASS <strong>AND</strong> SURVIVAL:<br />

ROLE OF ATP <strong>AND</strong> ROS PRODUCTION<br />

In type 2 diabetes, reduced β cell mass is a key feature associated with defects<br />

in insulin secretion. 61 The role of UCP2 in ATP production and ROS regulation<br />

has linked this protein not only to β cell function, but also to survival and<br />

proliferation. Increased ROS are thought to induce insulin resistance in numerous<br />

settings. 68 Undoubtedly, excessive ROS production is harmful to cell viability.<br />

However, moderately elevated ROS levels can also be essential for activating<br />

defensive signaling pathways to protect cells. 69<br />

For example, ROS can protect against ischemia–reperfusion injury in cardiomyocytes.<br />

2,69 The physiological role behind this phenomenon is that mitochondrial<br />

ROS is triggered to promote signaling pathways for gene transcription and<br />

cell growth in cells with low energy states. ROS may function as mediators to<br />

facilitate efficient energy transfer between mitochondria and ATPases in cells<br />

with high energy states. 70 Therefore, the beneficial effect of having lower levels<br />

of uncoupling activity in certain cell types like β cells under normal physiologic<br />

conditions may be to maintain a reasonable amount of ROS production. However,<br />

under pathological conditions when ROS production becomes excessive and<br />

harmful to cells, UCP2 may induce uncoupling of mitochondria in order to reduce<br />

ROS production.<br />

In support of this, evidence suggests that UCP2 controls mitochondrial ROS<br />

production which, in turn, confers neuroprotection, cardioprotection, and life span<br />

extension in Drosophila. 14,71,74 In the pancreatic β cell line INS-1, overexpression<br />

of UCP2 improved cell survival when the cells were treated with H 2O 2. 75 However,<br />

under chronic metabolic stress, even modestly increased levels of UCP2 protein<br />

may promote cell death in other types of tissues and cells including liver cells<br />

and Hela cells. 76–78<br />

With respect to cell proliferation, based on our study, UCP2 overexpression<br />

in MIN6 cells not only reduces mitochondrial ROS formation, but also increases<br />

proliferation under both starvation and nutritional excess conditions (unpublished


Role of Uncoupling Protein 2 in Pancreatic β-Cell Function 219<br />

data). Consistent with these observations, Ucp2 –/– mice had increased islet masses<br />

after 4.5 months of high fat diet feeding, suggesting that UCP2 may play a role<br />

in the regulation of β cell growth, probably by directly augmenting FFA-mediated<br />

ATP production. 51 The potential effect on cell growth is supported by the tumor<br />

formation that occurs upon deletion of UCP2 in association with increased NFκB<br />

action and oxidative stress. 79<br />

Islets of UCP2 –/– mice showed increased ROS levels when compared with<br />

islets from wild type (WT) mice. 80 In addition, FFAs also increased ROS production<br />

in isolated islets from WT and cultured pancreatic β cells, but not in<br />

islets from UCP2 –/– mice, 80,81 suggesting UCP2 contributes to the regulation of<br />

ROS production in β cells. However, data from our laboratory also suggest that<br />

despite the role of UCP2 in β cells, other factors may be involved in β cell<br />

regulation of intracellular ROS. Specifically, induction of ROS formation in vitro<br />

using menadione, ceramide, or cytokines in both Ucp2 –/– and WT islet cells failed<br />

to reveal any protective role of UCP2, perhaps because these mediators act beyond<br />

the UCP2 pathway (unpublished data). In addition, Ucp2 –/– and WT islets had<br />

equal sensitivities to apoptosis induced by menadione, ceramide, and cytokines;<br />

overexpression of UCP2 did not protect MIN6 cells from either cytokine- or<br />

menadione-mediated apoptosis. Thus, these data suggest that pancreatic β cells<br />

may have a specific regulatory signaling network to maintain their own levels of<br />

ROS, which may only be partially regulated by UCP2.<br />

CONCLUSION<br />

Through its fundamental role in the regulation of mitochondria ATP generation<br />

and ROS production, UCP2 is a key regulator of β cell function and survival.<br />

The pharmacological targeting of UCP2 in β cells presents a potential way to<br />

modulate insulin secretion and β cell proliferation. However, β cell-specific<br />

signaling pathways mediating the induction, activation, and effects of UCP2 need<br />

to be identified to enable precise manipulation without deleterious side effects.<br />

ACKNOWLEDGMENTS<br />

This work was funded by Grants MOP 12898 and MOP 43987 from the<br />

Canadian Institutes of Health Research (CIHR) to M.B.W. and C.B.C, respectively.<br />

M.B.W. is supported by an investigator award from CIHR. J.D. was<br />

supported by a fellowship award from the Canadian Diabetes Association. J.W.<br />

Joseph, V. Koshkin, A.V. Gyulkhandanyan, S.C. Lee, G.T. Karaman, C. Thorn,<br />

and G. Bikopoulos are gratefully acknowledged as contributors to the work<br />

presented here. We thank D. Yau, E. Allister, and N. Wijesekara for help with<br />

editing of the manuscript.


220 Oxidative Stress and Inflammatory Mechanisms<br />

REFERENCES<br />

1. Garlid KD, Jaburek M, Jezek P, and Varecha M. How do uncoupling proteins<br />

uncouple? Biochim. Biophys. Acta 2000; 1459: 383.<br />

2. Pain T et al. Opening of mitochondrial K(ATP) channels triggers the preconditioned<br />

state by generating free radicals. Circ. Res. 2000; 87: 460.<br />

3. Sohal RS and Weindruch R. Oxidative stress, caloric restriction, and aging. Science<br />

1996; 273: 59.<br />

4. Jezek P and Urbankova E. Specific sequence of motifs of mitochondrial uncoupling<br />

proteins. IUBMB Life 2000; 49: 63.<br />

5. Modriansky M, Murdza-Inglis DL, Patel HV, Freeman KB, and Garlid KD. Identification<br />

by site-directed mutagenesis of three arginines in uncoupling protein<br />

that are essential for nucleotide binding and inhibition. J. Biol. Chem. 1997; 272:<br />

24759.<br />

6. Jezek P and Freisleben HJ. Fatty acid binding site of the mitochondrial uncoupling<br />

protein: demonstration of its existence by EPR spectroscopy of 5-DOXYL-stearic<br />

acid. FEBS Lett. 1994; 343: 22.<br />

7. Rothwell NJ and Stock MJ. A role for brown adipose tissue in diet-induced<br />

thermogenesis. Nature 1979; 281: 31.<br />

8. Jacobsson A, Stadler U, Glotzer MA, and Kozak LP. Mitochondrial uncoupling<br />

protein from mouse brown fat: molecular cloning, genetic mapping, and mRNA<br />

expression. J. Biol. Chem. 1985; 260: 16250.<br />

9. Cannon B and Nedergaard J. Brown adipose tissue: function and physiological<br />

significance. Physiol. Rev. 2004; 84: 277.<br />

10. Jezek P and Garlid KD. Mammalian mitochondrial uncoupling proteins. Int. J.<br />

Biochem. Cell Biol. 1998; 30: 1163.<br />

11. Dridi S, Onagbesan O, Swennen Q, Buyse J, Decuypere E, and Taouis M. Gene<br />

expression, tissue distribution and potential physiological role of uncoupling protein<br />

in avian species. Comp. Biochem. Physiol. A Mol. Integr. Physiol. 2004; 139:<br />

273.<br />

12. Vidal-Puig AJ et al. Energy metabolism in uncoupling protein 3 gene knockout<br />

mice. J. Biol. Chem. 2000; 275: 16258.<br />

13. Saleh MC, Wheeler MB, and Chan CB. Uncoupling protein-2: evidence for its<br />

function as a metabolic regulator. Diabetologia 2002; 45: 174.<br />

14. Arsenijevic D et al. Disruption of the uncoupling protein-2 gene in mice reveals<br />

a role in immunity and reactive oxygen species production. Nat. Genet. 2000; 26:<br />

435.<br />

15. Chan CB, Saleh MC, Koshkin V, and Wheeler MB. Uncoupling protein 2 and<br />

islet function. Diabetes 2004; 53, Suppl 1: S136.<br />

16. Brand MD and Esteves TC. Physiological functions of the mitochondrial uncoupling<br />

proteins UCP2 and UCP3. Cell Metab. 2005; 2: 85.<br />

17. Andrews ZB, Diano S, and Horvath TL. Mitochondrial uncoupling proteins in the<br />

CNS: in support of function and survival. Nat. Rev. Neurosci. 2005; 6: 829.<br />

18. Krauss S, Zhang CY, and Lowell BB. The mitochondrial uncoupling-protein<br />

homologues. Nat. Rev. Mol. Cell Biol. 2005; 6: 248.<br />

19. Lowell BB and Shulman GI. Mitochondrial dysfunction and type 2 diabetes.<br />

Science 2005; 307: 384.


Role of Uncoupling Protein 2 in Pancreatic β-Cell Function 221<br />

20. Malaisse WJ, Hutton JC, Kawazu S, Herchuelz A, Valverde I, and Sener A. The<br />

stimulus-secretion coupling of glucose-induced insulin release. XXXV. The links<br />

between metabolic and cationic events. Diabetologia 1979; 16: 331.<br />

21. Freeman H, Shimomura K, Horner E, Cox RD, and Ashcroft FM. Nicotinamide<br />

nucleotide transhydrogenase: a key role in insulin secretion. Cell Metab. 2006; 3:<br />

35-45.<br />

22. Fleury C et al. Uncoupling protein-2: a novel gene linked to obesity and hyperinsulinemia.<br />

Nat. Genet. 1997; 15: 269.<br />

23. Gable DR, Stephens JW, Cooper JA, Miller GJ, and Humphries SE. Variation in<br />

the UCP2-UCP3 gene cluster predicts the development of type 2 diabetes in<br />

healthy middle-aged men. Diabetes 2006; 55: 1504.<br />

24. Bulotta A et al. The common -866G/A polymorphism in the promoter region of<br />

the UCP-2 gene is associated with reduced risk of type 2 diabetes in Caucasians<br />

from Italy. J. Clin. Endocrinol. Metab. 2005; 90: 1176.<br />

25. Vogler S et al. Association of a common polymorphism in the promoter of UCP2<br />

with susceptibility to multiple sclerosis. J. Mol. Med. 2005; 83: 806.<br />

26. Vogler S et al. Uncoupling protein 2 has protective function during experimental<br />

autoimmune encephalomyelitis. Am. J. Pathol. 2006; 168: 1570.<br />

27. Yamasaki H et al. Uncoupling protein 2 promoter polymorphism -866G/A affects<br />

peripheral nerve dysfunction in Japanese type 2 diabetic patients. Diabetes Care<br />

2006; 29: 888.<br />

28. Dalgaard LT and Pedersen O. Uncoupling proteins: functional characteristics and<br />

role in the pathogenesis of obesity and type 2 diabetes. Diabetologia 2001; 44: 946.<br />

29. Roduit R et al. Glucose down-regulates the expression of the peroxisome proliferator-activated<br />

receptor-alpha gene in the pancreatic beta cell. J. Biol. Chem.<br />

2000; 275: 35799.<br />

30. Patane G, Anello M, Piro S, Vigneri R, Purrello F, and Rabuazzo AM. Role of<br />

ATP production and uncoupling protein-2 in the insulin secretory defect induced<br />

by chronic exposure to high glucose or free fatty acids and effects of peroxisome<br />

proliferator-activated receptor-gamma inhibition. Diabetes 2002; 51: 2749.<br />

31. Medvedev AV et al. Regulation of the uncoupling protein-2 gene in INS-1 betacells<br />

by oleic acid. J. Biol. Chem. 2002; 277: 42639.<br />

32. Oberkofler H, Klein K, Felder TK, Krempler F, and Patsch W. Role of peroxisome<br />

proliferator-activated receptor-gamma coactivator-1alpha in the transcriptional<br />

regulation of the human uncoupling protein 2 gene in INS-1E cells. Endocrinology<br />

2006; 147: 966.<br />

33. Medvedev AV, Snedden SK, Raimbault S, Ricquier D, and Collins S. Transcriptional<br />

regulation of the mouse uncoupling protein-2 gene: double E-box motif is<br />

required for peroxisome proliferator-activated receptor-gamma-dependent activation.<br />

J. Biol. Chem. 2001; 276: 10817.<br />

34. Bordone L et al. Sirt1 regulates insulin secretion by repressing UCP2 in pancreatic<br />

beta cells. PLoS. Biol. 2006; 4: e31.<br />

35. Wolf G. Calorie restriction increases life span: a molecular mechanism. Nutr. Rev.<br />

2006; 64: 89.<br />

36. Lin SJ, Defossez PA, and Guarente L. Requirement of NAD and SIR2 for lifespan<br />

extension by calorie restriction in Saccharomyces cerevisiae. Science 2000;<br />

289: 2126.<br />

37. Cohen HY et al. Calorie restriction promotes mammalian cell survival by inducing<br />

the SIRT1 deacetylase. Science 2004; 305: 390.


222 Oxidative Stress and Inflammatory Mechanisms<br />

38. Motta MC et al. Mammalian SIRT1 represses forkhead transcription factors. Cell<br />

2004; 116: 551.<br />

39. Moynihan KA et al. Increased dosage of mammalian Sir2 in pancreatic beta cells<br />

enhances glucose-stimulated insulin secretion in mice. Cell Metab. 2005; 2: 105.<br />

40. Nemoto S, Fergusson MM, and Finkel T. Nutrient availability regulates SIRT1<br />

through a forkhead-dependent pathway. Science 2004; 306: 2105.<br />

41. Joseph JW et al. Uncoupling protein 2 knockout mice have enhanced insulin<br />

secretory capacity after a high-fat diet. Diabetes 2002; 51: 3211.<br />

42. Zhang CY et al. Uncoupling protein-2 negatively regulates insulin secretion and<br />

is a major link between obesity, beta cell dysfunction, and type 2 diabetes. Cell<br />

2001; 105: 745.<br />

43. Samec S, Seydoux J, and Dulloo AG. Post-starvation gene expression of skeletal<br />

muscle uncoupling protein 2 and uncoupling protein 3 in response to dietary fat<br />

levels and fatty acid composition: a link with insulin resistance. Diabetes 1999;<br />

48: 436.<br />

44. Li LX, Yoshikawa H, Egeberg KW, and Grill V. Interleukin-1-beta swiftly downregulates<br />

UCP-2 mRNA in beta-cells by mechanisms not directly coupled to<br />

toxicity. Cytokine 2003; 23: 101.<br />

45. Stylianou E and Saklatvala J. Interleukin-1. Int. J. Biochem. Cell Biol. 1998; 30:<br />

1075.<br />

46. Hohmeier HE, Tran VV, Chen G, Gasa R, and Newgard CB. Inflammatory mechanisms<br />

in diabetes: lessons from the beta cell. Int. J. Obes. Relat. Metab. Disord.<br />

2003; 27: S12.<br />

47. Maedler K et al. Glucose-induced beta cell production of IL-1-beta contributes to<br />

glucotoxicity in human pancreatic islets. J. Clin. Invest. 2002; 110: 851.<br />

48. Maedler K et al. Leptin modulates beta cell expression of IL-1 receptor antagonist<br />

and release of IL-1-beta in human islets. Proc. Natl. Acad. Sci. USA 2004; 101:<br />

8138.<br />

49. Baetens D, Malaisse-Lagae F, Perrelet A, and Orci L. Endocrine pancreas: threedimensional<br />

reconstruction shows two types of islets of Langerhans. Science 1979;<br />

206: 1323.<br />

50. Koster JC, Permutt MA, and Nichols CG. Diabetes and insulin secretion: the ATPsensitive<br />

K + channel (K ATP) connection. Diabetes 2005; 54: 3065.<br />

51. Joseph JW et al. Uncoupling protein 2 knockout mice have enhanced insulin<br />

secretory capacity after a high-fat diet. Diabetes 2002; 51: 3211.<br />

52. Chan CB et al. Increased uncoupling protein-2 levels in beta cells are associated<br />

with impaired glucose-stimulated insulin secretion: mechanism of action. Diabetes<br />

2001; 50: 1302.<br />

53. Wang MY et al. Adenovirus-mediated overexpression of uncoupling protein-2 in<br />

pancreatic islets of Zucker diabetic rats increases oxidative activity and improves<br />

beta-cell function. Diabetes 1999; 48: 1020.<br />

53a. Saleh MC, Wheeler MB, and Chan CB. Endogenous islet uncoupling protein-2<br />

expression and loss of glucose homeostasis in ob/ob mice. J. Endocrinol. 2006;<br />

190: 657.<br />

54. Briaud I, Kelpe CL, Johnson LM, Tran PO, and Poitout V. Differential effects of<br />

hyperlipidemia on insulin secretion in islets of Langerhans from hyperglycemic<br />

versus normoglycemic rats. Diabetes 2002; 51: 662.<br />

55. Kassis N et al. Correlation between pancreatic islet uncoupling protein-2 (UCP2)<br />

mRNA concentration and insulin status in rats. Int. J. Exp. Diab. Res. 2000; 1: 185.


Role of Uncoupling Protein 2 in Pancreatic β-Cell Function 223<br />

56. Takahashi A et al. Transgenic mice overexpressing nuclear SREBP-1c in pancreatic<br />

beta-cells. Diabetes 2005; 54: 492.<br />

57. Asghar Z, Yau D, Chan F, Leroith D, Chan CB, and Wheeler MB. Insulin resistance<br />

causes increased beta-cell mass but defective glucose-stimulated insulin secretion<br />

in a murine model of type 2 diabetes. Diabetologia 2006; 49: 90.<br />

58. Pecqueur C et al. Uncoupling protein 2, in vivo distribution, induction upon<br />

oxidative stress, and evidence for translational regulation. J. Biol. Chem. 2001;<br />

276: 8705.<br />

59. Jarmuszkiewicz W et al. Redox state of endogenous coenzyme q modulates the<br />

inhibition of linoleic acid-induced uncoupling by guanosine triphosphate in isolated<br />

skeletal muscle mitochondria. J. Bioenerg. Biomembr. 2004; 36: 4932.<br />

60. Koshkin V, Wang X, Scherer PE, Chan CB, and Wheeler MB. Mitochondrial<br />

functional state in clonal pancreatic beta-cells exposed to free fatty acids. J. Biol.<br />

Chem. 2003; 278: 19709.<br />

61. Hirabara SM et al. Acute effect of fatty acids on metabolism and mitochondrial<br />

coupling in skeletal muscle. Biochim. Biophys. Acta 2006; 1757: 57.<br />

62. Krauss S, Zhang CY, and Lowell BB. A significant portion of mitochondrial proton<br />

leak in intact thymocytes depends on expression of UCP2. Proc. Natl. Acad. Sci.<br />

USA 2002; 99: 118.<br />

63. Krauss S et al. Superoxide-mediated activation of uncoupling protein 2 causes<br />

pancreatic beta cell dysfunction. J. Clin. Invest. 2003; 112: 1831.<br />

64. Echtay KS et al. Superoxide activates mitochondrial uncoupling proteins. Nature<br />

2002; 415: 96.<br />

65. Haber EP, Procopio J, Carvalho CR, Carpinelli AR, Newsholme P, and Curi R.<br />

New insights into fatty acid modulation of pancreatic beta-cell function. Int. Rev.<br />

Cytol. 2006; 248: 1.<br />

66. Gromada J. The free fatty acid receptor GPR40 generates excitement in pancreatic<br />

beta-cells. Endocrinology 2006; 147: 672.<br />

67. Donath MY and Halban PA. Decreased beta-cell mass in diabetes: significance,<br />

mechanisms and therapeutic implications. Diabetologia 2004; 47: 581.<br />

68. Houstis N, Rosen ED, and Lander ES. Reactive oxygen species have a causal role<br />

in multiple forms of insulin resistance. Nature 2006; 440: 944.<br />

69. Sen CK and Packer L. Antioxidant and redox regulation of gene transcription.<br />

FASEB J. 1996; 10: 709.<br />

70. Garlid KD, Dos SP, Xie ZJ, Costa AD, and Paucek P. Mitochondrial potassium<br />

transport: role of the mitochondrial ATP-sensitive K(+) channel in cardiac function<br />

and cardioprotection. Biochim. Biophys. Acta 2003; 1606: 1.<br />

71. Bai Y et al. Persistent nuclear factor-kappa B activation in Ucp2–/– mice leads to<br />

enhanced nitric oxide and inflammatory cytokine production. J. Biol. Chem. 2005;<br />

280: 19062.<br />

72. Mattiasson G et al. Uncoupling protein-2 prevents neuronal death and diminishes<br />

brain dysfunction after stroke and brain trauma. Nat. Med. 2003; 9: 1062.<br />

73. Teshima Y, Akao M, Jones SP, and Marban E. Uncoupling protein-2 overexpression<br />

inhibits mitochondrial death pathway in cardiomyocytes. Circ. Res. 2003;<br />

93: 192.<br />

74. Fridell YW, Sanchez-Blanco A, Silvia BA, and Helfand SL. Targeted expression<br />

of the human uncoupling protein 2 (hUCP2) to adult neurons extends life span in<br />

the fly. Cell Metab. 2005; 1: 145.


224 Oxidative Stress and Inflammatory Mechanisms<br />

75. Li LX, Skorpen F, Egeberg K, Jorgensen IH, and Grill V. Uncoupling protein-2<br />

participates in cellular defense against oxidative stress in clonal beta-cells. Biochem.<br />

Biophys. Res. Commun. 2001; 282: 273.<br />

76. Chavin KD et al. Obesity induces expression of uncoupling protein-2 in hepatocytes<br />

and promotes liver ATP depletion. J. Biol. Chem. 1999; 274: 5692.<br />

77. Gnaiger E, Mendez G, and Hand SC. High phosphorylation efficiency and depression<br />

of uncoupled respiration in mitochondria under hypoxia. Proc. Natl. Acad.<br />

Sci. USA 2000; 97: 11080.<br />

78. Mills EM, Xu D, Fergusson MM, Combs CA, Xu Y, and Finkel T. Regulation of<br />

cellular oncosis by uncoupling protein 2. J. Biol. Chem. 2002; 277: 27385.<br />

79. Derdak Z et al. Enhanced colon tumor induction in uncoupling protein-2 deficient<br />

mice is associated with NF-kappa-B activation and oxidative stress. Carcinogenesis<br />

2006; 27: 956.<br />

80. Joseph JW et al. Free fatty acid-induced beta-cell defects are dependent on uncoupling<br />

protein 2 expression. J. Biol. Chem. 2004; 279: 51049.<br />

81. Koshkin V, Wang X, Scherer PE, Chan CB, and Wheeler MB. Mitochondrial<br />

functional state in clonal pancreatic beta-cells exposed to free fatty acids. J. Biol.<br />

Chem. 2003; 278: 19709.


Section II<br />

Influence of Dietary Factors,<br />

Micronutrients, and<br />

Metabolism


14<br />

CONTENTS<br />

Nutritional Modulation<br />

of Inflammation in<br />

Metabolic Syndrome<br />

Uma Singh, Sridevi Devaraj, and Ishwarlal Jialal<br />

Abstract .............................................................................................................227<br />

Metabolic Syndrome.........................................................................................228<br />

Metabolic Syndrome and Cardiovascular Disease...........................................228<br />

Metabolic Syndrome and Diabetes...................................................................229<br />

Inflammation, Metabolic Syndrome, and Acute Phase Proteins......................229<br />

Pro-Inflammatory Cytokines, Monocytes, and Metabolic Syndrome..............230<br />

Adipose Tissue and Inflammation ....................................................................231<br />

Therapeutic Modulation of Inflammation in Metabolic Syndrome.................232<br />

Weight Loss, Hypocaloric Diets, Inflammation, and Metabolic Syndrome....232<br />

Pharmacological Therapies with Potential to Prevent or Treat<br />

Metabolic Syndrome ...............................................................................237<br />

Conclusion.........................................................................................................238<br />

References .........................................................................................................238<br />

ABSTRACT<br />

Metabolic syndrome (MetS) is a disorder composed of central adiposity, dyslipidemia,<br />

abnormal glucose tolerance, and hypertension. It confers an increased<br />

risk for diabetes and cardiovascular disease (CVD). Inflammation plays a pivotal<br />

role in atherosclerosis and is involved in abnormalities associated with MetS such<br />

as insulin resistance (IR) and adiposity. The various biomarkers of inflammation<br />

such as inflammatory cytokines (TNF-α, IL-6, and IL-1β), chemokines (MCP-1<br />

and IL-8), and C-reactive protein (CRP) are increased in obesity and correlate<br />

with IR and CVD. IR is also associated with endothelial dysfunction (ED).<br />

Inflammation, IR, and ED amplify the cascade of metabolic and vascular<br />

derangements. The etiology of this syndrome is largely unknown but presumably<br />

represents a complex interaction of genetic and environmental factors. Since MetS<br />

is associated with chronic low grade inflammation, strategies are being explored<br />

227


228 Oxidative Stress and Inflammatory Mechanisms<br />

to ameliorate its pro-inflammatory status. MetS has been identified also as a target<br />

for diet therapy to reduce risk of CVD. Weight loss appears to be the best modality<br />

to reduce inflammation. Intervention trials convincingly demonstrate that weight<br />

loss and/or increased physical activity reduce biomarkers of inflammation such<br />

as CRP and IL-6. Thus, therapeutic lifestyle change (TLC) is strongly suggested<br />

for MetS subjects both for weight reduction and also to reduce the inflammation<br />

associated with this syndrome. Much more research is needed to define the roles<br />

of individual dietary factors on the biomarkers of inflammation and the mechanisms<br />

of the anti-inflammatory effects of weight loss in MetS.<br />

METABOLIC SYNDROME<br />

MetS comprises a cluster of abnormalities with insulin resistance and adiposity<br />

as its central features. 1–3 Five diagnostic criteria were identified by the Adult<br />

Treatment Panel III in the executive summary of a report of the National Cholesterol<br />

Education Program (NCEP). The presence of any three of the five features<br />

is considered sufficient to diagnose MetS. 4 These include central obesity, high<br />

triglycerides, low HDL, hypertension, and impaired fasting glucose. Using the<br />

NCEP definition on a representative sample of 8814 men and women from the<br />

United States, the age-adjusted prevalence of MetS was 24% in men and 23.4%<br />

in women. 5 Applying this figure to the U.S. population in the year 2000 means<br />

about 47 million residents have MetS.<br />

Recently, a worldwide consensus for the MetS defined by the International<br />

Diabetes Federation (IDF) includes central obesity as a main component and the<br />

other factors such as hypertension, dyslipidemia, and impaired fasting glucose<br />

as other components, all of which lead to insulin resistance and independently<br />

predict cardiovascular events. 6 The National Heart, Lung, and Blood Institute/American<br />

Heart Association statement defining MetS has just been published.<br />

It encompasses the recent IDF guidelines, but includes central obesity, dyslipidemia,<br />

hypertension, and impaired fasting glucose as the features. 7 This chapter<br />

will focus on biomarkers of inflammation in relation to MetS and its nutritional<br />

modulation. Particular emphasis will be given to nutritional insights derived<br />

from therapeutic interventions with diet and exercise in subjects with metabolic<br />

abnormalities.<br />

METABOLIC SYNDROME <strong>AND</strong> CARDIOVASCULAR <strong>DISEASE</strong><br />

Subjects with MetS have increased burdens of cardiovascular disease. In the<br />

Kuopio Ischemic Heart Disease Study, Lakka et al. 8 convincingly showed that<br />

men with MetS, even in the absence of baseline CAD or diabetes, had significantly<br />

increased mortality from CAD. In the Botnia Study, 9 MetS was defined as the<br />

presence of at least two of the following risk factors: obesity, hypertension,<br />

dyslipidemia, and microalbuminuria. Cardiovascular mortality was assessed in<br />

3606 subjects with a median follow-up of 6.9 years. In women and men,


Nutritional Modulation of Inflammation in Metabolic Syndrome 229<br />

respectively, MetS was seen in 10 and 15% of subjects with normal glucose<br />

tolerance (NGT), 42 and 64% of those with impaired fasting glucose/impaired<br />

glucose tolerance (IFG/IGT), and 78 and 84% of those with type 2 diabetes<br />

mellitus. The risk for coronary heart disease and stroke was increased three-fold<br />

in subjects with MetS (p


230 Oxidative Stress and Inflammatory Mechanisms<br />

markers and several features of MetS by conducting Z-score analyses in 107 nondiabetic<br />

subjects. CRP levels were shown to be strongly associated with insulin<br />

resistance calculated from the HOMA model, blood pressure, low HDL, triglycerides,<br />

and levels of the IL-6 and TNF pro-inflammatory cytokines. Body mass<br />

index (BMI) and insulin resistance were the strongest determinants of the inflammatory<br />

state.<br />

Festa et al. 15 who participated in the Insulin Resistance and Atherosclerosis<br />

Study (IRAS) showed that hsCRP was positively correlated with BMI, waist<br />

circumference, and other individual risk factors of MetS. Multivariate linear<br />

regression models reported the strongest correlation of CRP with BMI, systolic<br />

blood pressure, and insulin sensitivity index. All these studies record the universal<br />

observation that CRP is elevated in subjects with features of MetS and indicate<br />

a linear relationship between CRP levels and number of MetS features. Furthermore,<br />

the strongest associations are observed between CRP levels and central<br />

adiposity and insulin resistance.<br />

PRO-INFLAMMATORY CYTOKINES, MONOCYTES, <strong>AND</strong><br />

METABOLIC SYNDROME<br />

Cytokines, in particular IL-1, TNF, and IL-6, are the main inducers of the acute<br />

phase response (APR). 21 Several lines of evidence indicate that IL-6 is the central<br />

mediator of the inflammatory response and promotes insulin resistance. 22 In<br />

addition, IL-6 administered to humans subcutaneously induces an acute inflammatory<br />

response. 23 In fact, IL-6 is believed to be the main driver of CRP release<br />

from hepatocytes. 24<br />

A very tight correlation exists between IL-6 and CRP; such a strong correlation<br />

does not exist for other cytokines. Furthermore, IL-6 and hsCRP are<br />

associated with visceral adiposity and 30% of circulating IL-6 is derived from<br />

human adipose tissue. 25 Moreover, baseline IL-6 levels independently predict<br />

future CVD. 26 Pickup et al. 19 demonstrated increased levels of IL-6 in subjects<br />

with more than two features of MetS. Our group also showed that monocyte<br />

release of IL-6 in type 2 diabetes is significantly increased when compared to<br />

non-diabetic controls. 16 Several studies have shown that TNF secreted by monocytes–macrophages,<br />

endothelial cells, and also to a large extent by adipose tissue,<br />

is an important regulator of insulin sensitivity. 27<br />

Elevated IL-18 level has been recently reported as an independent risk predictor<br />

for MetS in the absence of a history of type 2 diabetes in a large communitybased<br />

population sample. 28 IL-18 functions as a pleiotropic pro-inflammatory<br />

cytokine and stimulates the production of TNF and secondarily IL-6. IL-18 may<br />

form a link between MetS and atherosclerosis because it is highly expressed in<br />

atherosclerotic plaques, and its role in plaque destabilization has been suggested. 29<br />

Additionally, data from the recently reported MONICA/KORA study show that<br />

elevated levels of IL-18 are associated with considerably increased risks of type<br />

2 diabetes. 30


Nutritional Modulation of Inflammation in Metabolic Syndrome 231<br />

Chemokines also play an important role in leukocyte trafficking. Circulating<br />

levels of chemokines have been shown to increase inflammatory processes including<br />

obesity-related pathologies. Kim et al. 31 reported that circulating levels of<br />

MCP-1 and IL-8 are related to obesity-related parameters such as BMI, waist<br />

circumference, CRP, IL-6, HOMA, and low HDL-cholesterol. These findings<br />

suggest that the circulating MCP-1 and/or IL-8 may be a potential candidate<br />

linking obesity with obesity-related metabolic complications such as atherosclerosis<br />

and diabetes.<br />

Since low grade inflammation is a feature of MetS, it is important to state<br />

that the pro-inflammatory status appears to result from an imbalance between<br />

pro- and anti-inflammatory cytokines. Importantly, in the Leiden 85-Plus Study,<br />

subjects who developed diabetes and had more than three MetS features had<br />

lower levels of LPS-stimulated whole blood release of IL-10 (a potent antiinflammatory<br />

cytokine) than those who did not. 32 Esposito et al. 33 also reported<br />

that in both obese and non-obese women, IL-10 levels were significantly lower<br />

in women with MetS.<br />

ADIPOSE TISSUE <strong>AND</strong> INFLAMMATION<br />

Changes in adipose tissue (AT) mass are associated with altered endocrine and<br />

metabolic functions of the adipose tissue. 34 Most importantly, adipose tissue is a<br />

depot of inflammatory molecules, collectively called the adipocytokines or adipokines,<br />

including tumor necrosis factor (TNF), angiotensinogen, PAI-1, leptin,<br />

and adiponectin. Unlike adipose tissues of lean individuals, adipose tissues of<br />

obese subjects secreted increased amounts of TNF, IL-6, iNOS, TGF-β, CRP,<br />

sICAM, MCP-1, procoagulant proteins such as PAI-1, and tissue factor. 35 Elevated<br />

PAI-1 levels, the principal inhibitors of fibrinolysis, have been reported in many<br />

clinical and population studies in obese subjects and correlate with abdominal<br />

patterns of obesity 36 and other components of MetS. Furthermore, it appears to<br />

be even a better predictor of the MetS status than CRP.<br />

A large body of evidence documents low levels of adiponectin in subjects<br />

with MetS. 37 Adiponectin belongs to the soluble defense collagen family and has<br />

been shown to suppress expression of adhesion molecules by endothelial cells,<br />

lipid accumulation, TNF release from monocytes, and vascular smooth muscle<br />

cell proliferation; it also reduces atherosclerotic vascular lesions in vivo. Adiponectin<br />

is an atypical adipokine because in contrast to the dramatic increase in<br />

plasma levels of all other adipokines, the circulating concentration of adiponectin<br />

is paradoxically decreased in obesity.<br />

A strong correlation between low levels of adiponectin and increased insulin<br />

resistance is well established in humans. 38 Adiponectin also improves insulin<br />

resistance in vivo. Among Asian Indians, plasma levels of adiponectin in subjects<br />

with IGT are strong predictors of the development of diabetes. 39 In a step-wise<br />

regression model of a study examining the association between CRP and adiponectin<br />

levels, hsCRP was independently associated inversely with levels of


232 Oxidative Stress and Inflammatory Mechanisms<br />

TABLE 14.1<br />

Metabolic Syndrome: Evidence for a<br />

Pro-Inflammatory State<br />

↑ Acute phase proteins (HsCRP and SAA)<br />

↓ Adiponectin<br />

↑ Pro-inflammatory status (IL-6, IL-18, and TNF-α)<br />

↑ Leptin<br />

↑ Chemokines (MCP-1 and IL-8)<br />

↓ Anti-inflammatory cytokines (IL-10)<br />

adiponectin and positively with levels of leptin. 40 Adiponectin exerts anti-atherogenic<br />

properties by suppressing the endothelial inflammatory response, inhibiting<br />

TNF and monocyte adhesion, and suppresses transformation of macrophages to<br />

foam cells.<br />

THERAPEUTIC MODULATION OF INFLAMMATION IN<br />

METABOLIC SYNDROME<br />

Inflammation can be reduced by a variety of approaches including diet, exercise,<br />

and pharmacotherapy (statins, PPAR-α and -γ agonists, and the endocannabinoid<br />

receptor-1 blocker, rimonabant). As reviewed previously 41 and summarized in<br />

Table 14.1, people with MetS typically manifest pro-thrombotic and pro-inflammatory<br />

states. Therapeutic lifestyle changes (TLCs), including diet and exercise,<br />

serve as cornerstone therapies for the treatment of MetS. Thus, the first step in<br />

reducing the excess cardiovascular risk associated with MetS is the adoption of<br />

a healthier lifestyle (particularly reducing body weight and increasing physical<br />

activity).<br />

WEIGHT LOSS, HYPOCALORIC DIETS, INFLAMMATION, <strong>AND</strong><br />

METABOLIC SYNDROME<br />

Weight losses ranging from 3 to 15 kg achieved through different diet programs<br />

(low fat, high protein, or hypocaloric diets) result in concomitant reductions of<br />

CRP levels by 7 to 48% as shown in several studies. We recently reviewed the<br />

literature 42,43 and documented a significant positive correlation (r 2 = 0.8693; p<br />


Nutritional Modulation of Inflammation in Metabolic Syndrome 233<br />

changes on markers of systemic vascular inflammation and IR, they conducted<br />

a randomized single-blind trial in 120 premenopausal (aged 20 to 46 years) obese<br />

women (BMI ≥30 kg/m 2 ) without diabetes, hypertension, or hyperlipidemia. The<br />

intervention group (n = 60) adhered to a low energy Mediterranean-style diet<br />

(foods rich in complex carbohydrates, monounsaturated fat, and fiber; lower ratios<br />

of omega-6 to omega-3 fatty acids) and increased physical activity. The control<br />

group (n = 60) was given general information about healthy food choices and<br />

exercise. BMI decreased significantly more in the intervention group than in<br />

controls, as did serum concentrations of IL-6, IL-18, and CRP, while adiponectin<br />

levels increased significantly. In multivariate analyses, changes in FFA and adiponectin<br />

levels were independently associated with changes in insulin sensitivity<br />

The same group of authors 45 investigated the effects of Mediterranean-style<br />

diets on endothelial function and vascular inflammatory markers in 180 patients<br />

with MetS. The 99 men and 81 women with MetS, as defined by the ATP III,<br />

were randomized equally in the intervention and placebo groups for 2 years. The<br />

patients in the intervention group were instructed to follow a Mediterranean-style<br />

diet and received detailed advice about how to increase daily consumption of<br />

whole grains, fruits, vegetables, nuts, and olive oil. Patients in the control group<br />

followed a prudent diet (50 to 60% carbohydrates, 15 to 20% proteins, total fat<br />


234 Oxidative Stress and Inflammatory Mechanisms<br />

TABLE 14.2<br />

Randomized Clinical Trials Examining Effects of Lifestyle Interventions<br />

(Diet and Exercise) on Biomarkers of Inflammation<br />

Study<br />

Investigators,<br />

Duration,<br />

Reference Subjects<br />

Esposito et al.<br />

2003<br />

(2 years) 44<br />

Esposito et al.<br />

2004<br />

(2 years) 45<br />

Tchernof et al.<br />

2002<br />

(14 months) 46<br />

Orchard et al.<br />

2005<br />

(3.2 years) 48<br />

Seshadari et al.<br />

2004<br />

(6 months) 50<br />

Kopp et al<br />

2003<br />

(14 months<br />

after surgery) 51<br />

120 premenopausal<br />

obese women (20 to<br />

46 years, BMI ≥30<br />

kg/m 2 )<br />

180 MetS patients (99<br />

men and 81 women)<br />

61 obese,<br />

postmenopausal<br />

women (mean age<br />

56.4 years, BMI-35.6<br />

kg/m 2 )<br />

Participants had IGT;<br />

657 controls and 638<br />

subjected to<br />

intervention<br />

78 severely obese<br />

patients, 86% with<br />

diabetes and/or MetS<br />

37 morbidly obese<br />

patients (BMI = 49<br />

kg/m 2 )<br />

Lifestyle Intervention<br />

Type<br />

Mediterranean-style diet<br />

(rich in complex<br />

carbohydrates, MUSFAs<br />

and fiber, lower ratio of<br />

omega 6/omega 3) and<br />

increased physical activity<br />

versus control subjects<br />

informed of healthy food<br />

choices and exercise<br />

Mediterranean style diet<br />

versus prudent diet (50 to<br />

60% carbohydrates, 15 to<br />

20% proteins,


Nutritional Modulation of Inflammation in Metabolic Syndrome 235<br />

TABLE 14.2 (CONTINUED)<br />

Randomized Clinical Trials Examining Effects of Lifestyle Interventions<br />

(Diet and Exercise) on Biomarkers of Inflammation<br />

Study<br />

Investigators,<br />

Duration,<br />

Reference Subjects<br />

Troseid et al.<br />

2004<br />

(12 weeks) 54<br />

Troseid et al.<br />

2005<br />

(12 weeks) 55<br />

Roberts et al.<br />

2005<br />

(3 weeks) 56<br />

Brinkworth et<br />

al. 2004<br />

(12 weeks<br />

energy<br />

restriction + 4<br />

weeks energy<br />

balance + 52<br />

weeks with<br />

minimal<br />

professional<br />

support) 57<br />

15 MetS subjects in<br />

exercise (n = 9) and<br />

non-exercise (n = 6)<br />

groups<br />

Lifestyle Intervention<br />

Type<br />

Endurance exercise<br />

(walking or jogging on<br />

treadmill) and strength<br />

training 45 to 60 minutes<br />

three times weekly in<br />

training studio under<br />

supervision<br />

32 subjects with MetS 2 × 2 randomized factorial<br />

trial, physical exercise<br />

31 obese patients; 15<br />

with MetS<br />

58 (13 male, 45 female)<br />

obese non-diabetic<br />

subjects with<br />

hyperinsulinemia;<br />

mean age =50 years,<br />

BMI = 34 kg/m 2<br />

IGT = impaired glucose tolerance. APN = adiponectin.<br />

High fiber, low fat diet; 3week<br />

residential program<br />

(ad libitum food and daily<br />

aerobic exercise)<br />

Randomization to standard<br />

protein (15% protein, 55%<br />

carbohydrate) or high<br />

protein (30% protein, 40%<br />

carbohydrate) diet<br />

Changes in Biomarkers<br />

of Inflammation<br />

Significant reduction in<br />

MCP-1 and IL-8<br />

No change in CAMs<br />

Significant reductions in<br />

CRP, sICAM, sP-selectin,<br />

MIP-1, MMP-9;<br />

increased NO; significant<br />

reduction in monocyte<br />

adhesion and chemotactic<br />

activity<br />

Significant (


236 Oxidative Stress and Inflammatory Mechanisms<br />

reviewed the importance of moderate increases in physical activity along with a<br />

detailed and tailored Mediterranean-style diet to reduce the prevalence of MetS<br />

and associated cardiovascular risks through reducing systemic inflammation and<br />

endothelium dysfunction, particularly in patients who failed to lose weight.<br />

In obese women, hypocaloric diets have been associated with weight reduction<br />

as well as inflammation. In this context, Seshadari et al. 50 found overall<br />

favorable effects on inflammation in 78 severely obese subjects from a low<br />

carbohydrate diet (LCD) versus a conventional (fat- and calorie-restricted) diet<br />

for a period of 6 months; 86% of the subjects were either diabetic or had features<br />

of MetS. Overall, CRP levels decreased modestly in both diet groups. However,<br />

patients with high risk baseline levels (CRP >3 mg/dL, n = 48) experienced<br />

greater decreases in CRP on a low carbohydrate diet, independent of weight loss.<br />

Kopp et al. 51 examined the cross-sectional and longitudinal relationships of<br />

CRP, IL-6, and TNF with features of the IRS in 37 morbidly obese patients (BMI<br />

= 49 kg/m 2 ) with different stages of glucose tolerance before and 14 months after<br />

gastric surgery. Weight loss after gastric surgery induced a significant shift from<br />

diabetes (37 versus 3%) to IGT (40 versus 33%) and NGT (23 versus 64%).<br />

Concentrations of CRP and IL-6 decreased after weight loss, whereas serum<br />

levels of TNF-alpha remained unchanged. It was concluded that weight loss<br />

results in a marked decrease in circulating levels of inflammatory markers in<br />

association with a reversal of diabetes in morbidly obese individuals after gastroplastic<br />

surgery. However, long-term studies are needed to show whether this<br />

improvement in cardiovascular risk factors will eventually translate into a significant<br />

clinical benefit with regard to cardiovascular morbidity and mortality.<br />

Cellular adhesion molecules (CAMs) such as E-selectin, intercellular adhesion<br />

molecule-1 (ICAM-1), and vascular cell adhesion molecule-1 (VCAM-1)<br />

are involved in the rolling, adhesion, and extravasation of monocytes and T<br />

lymphocytes into the atherosclerotic plaque. Serum concentrations of CAMs are<br />

higher in patients with coronary artery disease than in healthy control subjects. 52<br />

Moreover, male participants in the Physician’s Health Study who had ICAM-1<br />

levels in the highest quartile are at greater cardiovascular risk than men in the<br />

lowest quartile. 53 In line with these findings, Troseid et al. 54 published a study<br />

conducted as an unmasked randomized 2 × 2 factorial trial involving an intensive<br />

exercise protocol of 12 weeks’ duration. In the combined exercise groups, significant<br />

reductions in MCP-1 and IL-8 as compared to the combined non-exercise<br />

groups were noted along with a significant reduction versus baseline for both<br />

chemokines. The changes in MCP-1 were significantly correlated to changes in<br />

visceral fat.<br />

However, the same group of investigators 55 recently reported a negative study<br />

with regard to the effect of physical exercise on serum levels of CAMs. These<br />

authors explored the possible role of adipose tissue in regulating serum levels of<br />

CAMs. No significant changes in CAMs were observed in the intervention group.<br />

On examination of the whole study population regardless of intervention, changes<br />

in serum E-selectin were significantly correlated to changes in body mass index,<br />

waist circumference, fasting glucose, and HbA1c, but not to changes in visceral


Nutritional Modulation of Inflammation in Metabolic Syndrome 237<br />

fat, subcutaneous fat, TNF, or adiponectin. Thus this study highlights changes in<br />

glycemic control and obesity rather than regional fat distribution to influence Eselectin<br />

levels in subjects with MetS.<br />

Additionally, Roberts et al. 56 recently reported the results of examining the<br />

effects of lifestyle modification on various key contributing factors to atherogenesis,<br />

including oxidative stress, inflammation, chemotaxis, and cell adhesion.<br />

Obese men (n = 31), 15 of whom had MetS, were placed on a high fiber, low fat<br />

diet in a 3-week residential program where food was provided ad libitum and<br />

daily aerobic exercise (45 to 60 min) was performed. After 3 weeks, significant<br />

reductions in BMI, CRP, sICAM-1, sP-selectin, MIP-1α, MMP-9, and biomarkers<br />

of oxidative stress with concomitant increases in NO production were noted.<br />

Additionally, both monocyte adhesion and monocyte chemotactic activity (MCA)<br />

significantly decreased. Nine of 15 subjects were no longer positive for MetS<br />

post-intervention. This study postulated the amelioration of CAD risk factor by<br />

intensive lifestyle modification in men with MetS factors prior to reversal of<br />

obesity.<br />

Comparative effects of two low fat diets differing in carbohydrate-to-protein<br />

ratio for biomarkers of CVD risk in obese subjects with hyperinsulinemia were<br />

examined. 57 Two groups totaling 58 obese, non-diabetic subjects with hyperinsulinemia<br />

(13 males and 45 females, mean age 50.2 years, mean BMIs of 34.0<br />

kg/m 2 , mean fasting insulin levels of 17.8 mU/l) were randomly assigned to either<br />

a standard protein (SP; 15% protein, 55% carbohydrate) or high protein (HP;<br />

30% protein, 40% carbohydrate) diet during 12 weeks of energy restriction<br />

(approximately 6.5 MJ/day) and 4 weeks of energy balance (approximately 8.3<br />

MJ/day). They were subsequently asked to maintain the same dietary pattern for<br />

the succeeding 52 weeks with minimal professional support. The measurements<br />

included fasting blood lipids, glucose, insulin, and CRP and sICAM-1 at baseline<br />

and at weeks 16 and 68. In total, 43 subjects completed the study with similar<br />

dropouts in each group. At week 68, there was net weight loss (SP –2.9 ± 3.6%;<br />

HP –4.1 ± 5.8%; p


238 Oxidative Stress and Inflammatory Mechanisms<br />

Chief among these are statins, insulin sensitizers such as metformin and<br />

thiazolidinediones (rosiglitazone and pioglitazone), fibrates, and a newly discovered<br />

drug named rimonabant. In various large prospective studies, statins have<br />

been shown to reduce hsCRP. 59 Rosiglitazone has also been shown to significantly<br />

reduce hsCRP levels, 60 suggesting that PPAR-γ agonists may reduce markers for<br />

subclinical inflammation to the levels seen in statin trials. Furthermore, therapy<br />

with fenofibrate, a fibric acid derivative, has been shown to lower IL-6 and hsCRP<br />

in borderline hyperlipidemic subjects. 61 In addition, the endocannabinoid system<br />

appears to play a key role in metabolism and weight gain. The investigational<br />

rimonabant is a cannabinoid receptor type 1 blocker that has been employed in<br />

trials involving more than 6500 patients. It has resulted in reduction in the<br />

prevalence of MetS as well as biomarkers of inflammation along with improved<br />

glycemic and lipid profiles. 58,62<br />

CONCLUSION<br />

It appears from the available literature that therapeutic lifestyle change remains<br />

the cornerstone in modulating inflammation and CVD. If lifestyle change is not<br />

sufficient, then drug therapies for abnormalities in individual risk factors are to<br />

be indicated. To date, we have insufficient evidence for primary use of drugs that<br />

target the underlying causes of MetS. Weight loss and hypocaloric diets are<br />

definitely associated with reduced inflammation and overall risk reduction of<br />

CVD.<br />

Also, most studies reported effects on CRP, whereas a few focused on proinflammatory<br />

cytokines such as IL-6, IL-18 or TNF-α and chemokines (IL-8 and<br />

MCP-1). Future research should focus on the roles of specific nutritional components<br />

in various diets on biomarkers of inflammation, as they may modulate<br />

inflammation through different mechanisms. Despite the fact that precise mechanisms<br />

have yet to be established, both diet and physical activity play pivotal<br />

roles in improving many factors associated with MetS including modulation of<br />

various adipocytokines. A more thorough understanding of the clustering of<br />

metabolic abnormalities and their underlying etiology will help to define diet and<br />

physical activity guidelines for preventing and treating MetS — an important<br />

aspect of CVD prevention.<br />

REFERENCES<br />

1. Reaven, G.M. (2005) The insulin resistance syndrome: definition and dietary<br />

approaches to treatment. Ann Rev Nutr 25, 391.<br />

2. Eckel, R.H., Grundy, S. M., and Zimmet, P.Z. (2005) The metabolic syndrome.<br />

Lancet 365, 1415.<br />

3. Haffner, S. and Cassells, H.B. (2003) Metabolic syndrome: a new risk factor of<br />

coronary heart disease? Diabetes Obes Metab 5, 359.


Nutritional Modulation of Inflammation in Metabolic Syndrome 239<br />

4. Expert Panel on Detection, Evaluation, and Treatment of High Blood Cholesterol<br />

in Adults. (2001) Executive Summary of Third Report of the National Cholesterol<br />

Education Program (NCEP) Expert Panel on Detection, Evaluation, and Treatment<br />

of High Blood Cholesterol in Adults (Adult Treatment Panel III). (2001) JAMA<br />

285, 2486.<br />

5. Ford, E.S., Giles, W.H., and Dietz, W.H. (2002) Prevalence of the metabolic<br />

syndrome among U.S. adults: findings from the third National Health and Nutrition<br />

Examination Survey. JAMA 287, 356.<br />

6. Alberti, K.G, Zimmet, P., and Shaw, J. (2005) IDF Epidemiology Task Force<br />

Consensus Group: the metabolic syndrome — a new worldwide definition. Lancet<br />

366, 1059.<br />

7. Grundy, S.M. et al. (2005) Diagnosis and management of the metabolic syndrome:<br />

an American Heart Association/National Heart, Lung, and Blood Institute scientific<br />

statement. Circulation 112, 2735.<br />

8. Lakka, H.M. et al. (2002) The metabolic syndrome and total and cardiovascular<br />

disease mortality in middle-aged men. JAMA 288, 2709.<br />

9. Isomaa, B. et al. (2001) Cardiovascular morbidity and mortality associated with<br />

the metabolic syndrome. Diabetes Care 24, 683.<br />

10. Alexander, C.M., Landsman, P.B., Teutsch, S.M., and Haffner, S.M. (2003) Third<br />

National Health and Nutrition Examination Survey (NHANES III); National Cholesterol<br />

Education Program (NCEP). NCEP-defined metabolic syndrome, diabetes,<br />

and prevalence of coronary heart disease among NHANES III participants<br />

age 50 years and older. Diabetes 52, 1210.<br />

11. Dekker, J.M. et al. (2005) Metabolic syndrome and 10-year cardiovascular disease<br />

risk in the Hoorn Study. Circulation 112, 666.<br />

12. Hanson, R.L, Imperatore, G., Bennett, P.H., and Knowler, W.C. (2002) Components<br />

of the “metabolic syndrome” and incidence of type 2 diabetes. Diabetes 51,<br />

3120.<br />

13. Wallace, A.M. et al. (2001) Plasma leptin and the risk of cardiovascular disease<br />

in the West of Scotland Coronary Prevention Study (WOSCOPS). Circulation<br />

104, 3052.<br />

14. Assmann, G., Nofer, J.R., and Schulte, H. (2004) Cardiovascular risk assessment<br />

in metabolic syndrome: view from PROCAM. Endocrinol Metab Clin North Am<br />

33, 377.<br />

15. Festa, A. et al. (2000) Chronic subclinical inflammation as part of the insulin<br />

resistance syndrome: the Insulin Resistance Atherosclerosis Study (IRAS). Circulation<br />

102, 42.<br />

16. Devaraj, S. and Jialal, I. (2000) Alpha-tocopherol supplementation decreases<br />

serum C-reactive protein and monocyte interleukin-6 levels in normal volunteers<br />

and type 2 diabetic patients. Free Radic Biol Med 29, 790.<br />

17. Ridker, P.M., Buring, J.E., Cook, N.R., and Rifai, N. (2003) CRP, the metabolic<br />

syndrome, and risk of incident cardiovascular events: an 8-year follow-up of 14719<br />

initially healthy American women. Circulation 107, 391.<br />

18. Ford, E.S. (2003) The metabolic syndrome and C-reactive protein, fibrinogen, and<br />

leukocyte count: findings from the Third National Health and Nutrition Examination<br />

Survey. Atherosclerosis 168, 351.<br />

19. Pickup, J.C., Mattock, M.B., Chusney, G.D., and Burt, D. (1997) NIDDM as a<br />

disease of the innate immune system: association of acute-phase reactants and<br />

interleukin-6 with metabolic syndrome X. Diabetologia 40, 1286.


240 Oxidative Stress and Inflammatory Mechanisms<br />

20. Yudkin, J.S., Kumari, M, Humphries, S.E., and Mohamed-Ali, V. (2000) Inflammation,<br />

obesity, stress and coronary heart disease: is interleukin-6 the link? Atherosclerosis<br />

148, 209.<br />

21. Baumann, H. and Goldie, J. (1994) The acute phase response. Immunol Today 15,<br />

74.<br />

22. Heinrich, P.C., Castell, J.V., and Andus T. (1990) Interleukin-6 and the acute phase<br />

response. Biochem J 265, 621.<br />

23. Banks, R.E. et al. (1995) The acute phase protein response in patients receiving<br />

subcutaneous IL-6. Clin Exp Immunol 102, 217.<br />

24. Li, S.P. and Goldman, N.D. (1996) Regulation of human C-reactive protein gene<br />

expression by two synergistic IL-6 responsive elements. Biochemistry 35, 9060.<br />

25. Aldhahi, W. and Hamdy, O. (2003) Adipokines, inflammation, and the endothelium<br />

in diabetes. Curr Diab Rep 3, 293.<br />

26. Ridker, P.M., Hennekens, C.H., Buring, J.E., and Rifai, N. (2000) C-reactive<br />

protein and other markers of inflammation in the prediction of cardiovascular<br />

disease in women. New Engl J Med 342, 836.<br />

27. Hotamisligil, G.S. and Spiegelman, B.M. (1994) Tumor necrosis factor-alpha: a<br />

key component of the obesity-diabetes link. Diabetes 43, 1271.<br />

28. Hung, J., McQuillan, B.M., Chapman, C.M., Thompson, P.L., and Beilby, J.P.<br />

(2005) Elevated interleukin-18 levels are associated with the metabolic syndrome<br />

independent of obesity and IR. Arterioscler Thromb Vasc Biol 25, 1268.<br />

29. SoRelle, R. (2003) Interleukin-18 predicts coronary events. Circulation 108,<br />

e9051.<br />

30. Thorand, B. et al. (2005) Elevated levels of interleukin-18 predict the development<br />

of type 2 diabetes: results from the MONICA/KORA Augsburg Study, 1984–2002.<br />

Diabetes 54, 2932.<br />

31. Kim, C.S., Park, H.S., Kawada, T., Kim, J.H., and Erickson, K.L. (2006) Circulating<br />

levels of MCP-1 and IL-8 are elevated in human obese subjects and associated<br />

with obesity-related parameters. Int J Obes [epub ahead of print].<br />

32. van Exel, E., Gussekloo, J., and de Craen, A.J. (2002) Low production capacity<br />

of interleukin-10 associates with the metabolic syndrome and type 2 diabetes: the<br />

Leiden 85-Plus Study. Diabetes 51, 1088.<br />

33. Esposito, K., Pontillo, A., and Giugliano, F. (2003). Association of low interleukin-<br />

10 levels with the metabolic syndrome in obese women. J Clin Endocrinol Metab<br />

88, 1055.<br />

34. Lau, D.C., Dhillon, B., Yan, H., Szmitko, P.E., and Verma S. (2005) Adipokines:<br />

molecular links between obesity and atherosclerosis. Am J Physiol Heart Circ<br />

Physiol 288, H2031.<br />

35. Rajala, M.W. and Scherer, P.E. (2003) Minireview: the adipocyte at the crossroads<br />

of energy homeostasis, inflammation, and atherosclerosis. Endocrinology 144,<br />

3765.<br />

36. Sakkinen, P.A, Wahl, P., Cushman, M., Lewis, M.R., and Tracy, R.P. (2000)<br />

Clustering of procoagulation, inflammation, and fibrinolysis variables with metabolic<br />

factors in insulin resistance syndrome. Am J Epidemiol 152, 897.<br />

37. Kondo, H., Shimomura, I., and Matsukawa, Y. (2002) Association of adiponectin<br />

mutation with type 2 diabetes: a candidate gene for the insulin resistance syndrome.<br />

Diabetes 51, 2325.


Nutritional Modulation of Inflammation in Metabolic Syndrome 241<br />

38. Weyer, C., Funahashi, T., and Tanaka, S. (2001) Hypoadiponectinemia in obesity<br />

and type 2 diabetes: close association with insulin resistance and hyperinsulinemia.<br />

J Clin Endocrinol Metab 86, 1930.<br />

39. Snehalatha, C., Mukesh, B., and Simon, M. (2003) Plasma adiponectin is an<br />

independent predictor of T2DM in Asian Indians. Diabetes Care 26, 3226.<br />

40. Ouchi, N. et al. (2003) Reciprocal association of C-reactive protein with adiponectin<br />

in blood stream and adipose tissue. Circulation 107, 671.<br />

41. Devaraj, S., Rosenson, R.S., and Jialal, I. (2004) Metabolic syndrome: an appraisal<br />

of the pro-inflammatory and procoagulant status. Endocrinol Metab Clin North<br />

Am 33, 431.<br />

42. Basu, A., Devaraj, S., and Jialal, I. (2006) Dietary factors that promote or retard<br />

inflammation. Arterioscler Thromb Vasc Biol 26, 995.<br />

43. Dietrich, M. and Jialal, I. (2005) The effect of weight loss on a stable biomarker<br />

of inflammation, C-reactive protein. Nutr Rev 63, 22.<br />

44. Esposito, K. et al. (2003) Effect of weight loss and lifestyle changes on vascular<br />

inflammatory markers in obese women: a randomized trial. JAMA 289, 1799.<br />

45. Esposito, K. et al. (2004) Effect of a Mediterranean-style diet on endothelial<br />

dysfunction and markers of vascular inflammation in metabolic syndrome: a<br />

randomized trial. JAMA 292, 1440.<br />

46. Tchernof, A., Nolan, A., Sites, C.K., Ades, P.A., and Poehlman, E.T. (2002) Weight<br />

loss reduces C-reactive protein levels in obese postmenopausal women. Circulation.<br />

105, 564.<br />

47. Diabetes Prevention Program (DPP) Research Group. (2002) The Diabetes Prevention<br />

Program (DPP): description of lifestyle intervention. Diabetes Care 25,<br />

2165.<br />

48. Orchard, T.J. et al. (2005) The effect of metformin and intensive lifestyle intervention<br />

on the metabolic syndrome: the Diabetes Prevention Program randomized<br />

trial. Ann Intern Med 142, 611.<br />

49. Meydani, M. (2005). A Mediterranean-style diet and metabolic syndrome. Nutr<br />

Rev 63, 312.<br />

50. Seshadri, P. et al. (2004) A randomized study comparing the effects of a lowcarbohydrate<br />

diet and a conventional diet on lipoprotein subfractions and Creactive<br />

protein levels in patients with severe obesity. Am J Med 117, 398.<br />

51. Kopp, H.P. et al. (2003) Impact of weight loss on inflammatory proteins and their<br />

association with the insulin resistance syndrome in morbidly obese patients. Arterioscler<br />

Thromb Vasc Biol 23, 1042.<br />

52. Haught, W.H., Mansour, M., and Rothlein, R. (1996) Alterations in circulating<br />

intercellular adhesion molecule-1 and L-selectin further evidence for chronic<br />

inflammation in ischemic heart disease. Am Heart J 132, 1.<br />

53. Ridker, P.M., Hennekens, H., Roitman-Johnson, M.B., Stampfer, J., and Allen, J.<br />

(1998) Plasma concentration of soluble intercellular adhesion molecule 1 and risks<br />

of future myocardial infarction in apparently healthy men, Lancet 351, 88.<br />

54. Troseid, M., Lappegard, K.T., and Claudi T. (2004) Exercise reduces plasma levels<br />

of chemokines MCP-1 and IL-8 in subjects with the metabolic syndrome. Eur<br />

Heart J 25, 349.<br />

55. Troseid, M., Lappegard, K.T., Mollnes, T.E., Arnesen, H., and Seljeflot, I. (2005)<br />

Changes in serum levels of E-selectin correlate to improved glycaemic control<br />

and reduced obesity in subjects with the metabolic syndrome. Scand J Clin Lab<br />

Invest 65, 283.


242 Oxidative Stress and Inflammatory Mechanisms<br />

56. Roberts, C.K. et al. (2005) Effect of a diet and exercise intervention on oxidative<br />

stress, inflammation, MMP-9 and monocyte chemotactic activity in men with<br />

metabolic syndrome factors. J Appl Physiol 100, 1657.<br />

57. Brinkworth, G.D., Noakes, M., Keogh, J.B., Wittert, G.A., and Clifton, P.M. (2004)<br />

Long-term effects of a high-protein, low-carbohydrate diet on weight control and<br />

cardiovascular risk markers in obese hyperinsulinemic subjects. Int J Obes Relat<br />

Metab Disord 28, 661.<br />

58. Davis, S.N. (2006) Contemporary strategies for managing cardiometabolic risk<br />

factors. J Manag Care Pharm 12, S4.<br />

59. Devaraj, S. and Jialal, I. (2004) Effects of statins on C-reactive protein: are all<br />

statins similar? In Statins: Understanding Clinical Use, Elsevier Science, Amsterdam,<br />

p. 189.<br />

60. Esposito, K. et al. (2006) Effect of rosiglitazone on endothelial function and<br />

inflammatory markers in patients with the metabolic syndrome. Diabetes Care<br />

29, 1071.<br />

61. Zambon, A., Gervois, P., Pauletto, P., Fruchart, J.C., and Staels, B. (2006) Modulation<br />

of hepatic inflammatory risk markers of cardiovascular diseases by PPARalpha<br />

activators: clinical and experimental evidence. Arterioscler Thromb Vasc<br />

Biol 26, 977.<br />

62. Despres, J.P., Golay, A., and Sjostrom, L. (2005) Rimonabant in Obesity-Lipids<br />

Study Group: effects of rimonabant on metabolic risk factors in overweight<br />

patients with dyslipidemia. New Engl J Med 353, 2121.


15<br />

CONTENTS<br />

Dietary Fatty Acids and<br />

Metabolic Syndrome<br />

Helen M. Roche<br />

Introduction: Dietary Fat Intake, Inflammation, and Metabolic<br />

Syndrome —Double Hit Hypothesis ......................................................243<br />

Dietary Fatty Acids, Inflammation, and Insulin Signalling:<br />

Cellular Perspective.................................................................................244<br />

Whole Body Metabolic Perspective: Evidence from Human<br />

Dietary Fatty Acid Intervention Studies .................................................246<br />

Future Perspectives ...........................................................................................248<br />

References .........................................................................................................248<br />

INTRODUCTION: DIETARY FAT INTAKE, INFLAMMATION, <strong>AND</strong><br />

METABOLIC SYNDROME —DOUBLE HIT HYPOTHESIS<br />

Nutrition is a key environmental factor that is particularly involved in the pathogenesis<br />

and progression of several polygenic, diet-related diseases. Metabolic<br />

syndrome is a very common condition, characterized by insulin resistance,<br />

abdominal obesity, dyslipidemia [increased triacylglycerol (TAG) or reduced high<br />

density lipoprotein (HDL) cholesterol concentrations], hypertension, and urinary<br />

microalbuminuria. 1–3 It often precedes type 2 diabetes mellitus (T2DM) and is<br />

associated with a greater risk of cardiovascular disease (CVD). 4,5<br />

The development of metabolic syndrome can be attributed to both genetic<br />

and environmental factors, as reviewed elsewhere. 6 Figure 15.1 illustrates that<br />

insulin resistance is the most important metabolic defect that leads to the development<br />

of the metabolic syndrome — a state that may be triggered by excessive<br />

metabolic stressors [high fat diet, obesity, elevated plasma non-esterified fatty<br />

acid (NEFA) levels, etc.]. 7,8<br />

The link between dietary fat, obesity, and T2DM is partly due to elevated<br />

fatty acid concentrations on systemic responsiveness to insulin, resulting in<br />

impaired insulin action in several peripheral tissues including the liver, skeletal<br />

muscles, and adipose tissues and dysregulated carbohydrate and lipid metabolism,<br />

which leads to a compensatory hyperinsulinemia. 9 Interestingly, insulin resistance<br />

243


244 Oxidative Stress and Inflammatory Mechanisms<br />

Progressive phenotype<br />

Metabolic<br />

stress<br />

Sensitive Genotype<br />

Obesity<br />

Insulin resistance<br />

Glucose intolerance<br />

Metabolic syndrome<br />

T 2 DM<br />

Inflammation<br />

FIGURE 15.1 Development and progression of metabolic syndrome. (From Roche, H.M.,<br />

Proc Nutr Soc 64, 23, 2005. With permission.)<br />

often co-exists with a subacute chronic pro-inflammatory state. 10–12 Indeed, recent<br />

research suggests that pro-inflammatory cytokines derived from adipose tissue<br />

play a key role in the development of insulin resistance. 13,14 The model in Figure<br />

15.1 suggests that an individual with a genetic predisposition to metabolic syndrome<br />

cannot sustain a “double hit” from metabolic and inflammatory stressors.<br />

Also, the pathway between obesity and insulin resistance toward the metabolic<br />

syndrome and T2DM represents a progressive phenotype. Therefore, attenuating<br />

the impacts of metabolic stressors through dietary fat modification to improve<br />

insulin sensitivity would be advantageous.<br />

DIETARY FATTY ACIDS, INFLAMMATION, <strong>AND</strong> INSULIN<br />

SIGNALLING: CELLULAR PERSPECTIVE<br />

Several lines of evidence suggest that the combination of excessive nutrientderived<br />

metabolic stressors and pro-inflammatory stressors plays an important<br />

role in the development of insulin resistance and metabolic syndrome. High fat<br />

diets and excessive adipose tissue TAG storage result in increased circulating<br />

plasma NEFA flux to peripheral tissues that promotes insulin resistance. Fatty<br />

acids induce insulin resistance through inhibition of insulin signaling. 15–17<br />

Evidence also indicates that saturated fatty acids (SFAs) may play a particular<br />

role in attenuating peripheral tissue responsiveness to insulin 3 while some polyunsaturated<br />

fatty acids (PUFAs) counteract this effect. Adipose tissue may be the<br />

source of insulin de-sensitization of pro-inflammatory molecules collectively<br />

known as adipocytokines or adipokines, which predispose to insulin resistance.<br />

13,14 Tumor necrosis factor (TNF)-α, interleukin (IL)-6, resistin, and adiponectin<br />

(acrp30) have all been shown to influence insulin sensitivity. 18–20 The<br />

insulin de-sensitizing effects of TNF-α are probably best characterized. TNF-α<br />

Molecular mechanisms


Dietary Fatty Acids and Metabolic Syndrome 245<br />

inhibits autophosphorylation of tyrosine residues of the insulin receptor, promotes<br />

serine phosphorylation of insulin receptor substrate (IRS)-1, and reduces transcription<br />

of key targets in the insulin signaling cascade, all of which impede<br />

transduction of the insulin signal. 21<br />

Knocking out TNF-α or the TNF-α receptor improves insulin resistance in<br />

animal models of obesity-induced insulin resistance. 22,23 Other groups have proposed<br />

that other components of the inflammatory response, IκB kinase-β (IKKβ)<br />

or c-Jun amino-terminal kinases (JNKs), are central to the interplay of dietary<br />

fatty acids, obesity, and insulin resistance. 24,25 In summary, these studies suggest<br />

that down-regulation of several components of the inflammatory response affords<br />

substantial protection from obesity-induced insulin resistance.<br />

Consistent with our hypothesis that dietary fatty acids can modulate the<br />

inflammatory profiles of adipocytes, two recent studies determined the effects of<br />

other fatty acids in vitro. The first showed that the palmitate SFA activated nuclear<br />

factor-κB (NF-κB) activity and induced TNF-α and IL-6 expression in 3T3-L1<br />

adipocytes. 26 The second extensive investigation demonstrated that a mixture of<br />

SFA and PUFA NEFA treatments impaired insulin signalling at multiple sites,<br />

decreased insulin-stimulated glucose transporter (GLUT)-4 translocation and glucose<br />

transport, and activated the stress/inflammatory kinase JNK pathway in 3T3-<br />

L1 adipocytes. 27 Thus it may be possible to attenuate the pro-inflammatory phenotype<br />

associated with obesity-induced insulin resistance by altering the composition<br />

of circulating NEFAs through dietary fatty acid modification.<br />

Therefore, in terms of manipulating dietary factors to attenuate the inflammatory<br />

response in adipose tissue to improve insulin sensitivity, the most obvious<br />

treatment is to reduce adipose tissue mass. Nevertheless the prevalence of obesity<br />

is increasing and due to poor compliance, current therapies are largely ineffective.<br />

Therefore other strategies to attenuate the impact of insulin resistance in the<br />

presence of obesity are required.<br />

Our group demonstrated that a sub-group of fatty acids known as conjugated<br />

linoleic acids and, in particular, the cis-9, trans-11 CLA isomer (c9, t11-CLA),<br />

may have the potential to improve lipid metabolism and insulin sensitivity within<br />

the context of obesity. 28,29 This effect was ascribed to differential sterol regulatory<br />

element-binding protein (SREBP)-1c gene expression, a key regulatory transcription<br />

factor involved in lipogenesis and glucose metabolism. Feeding a c9, t11-<br />

CLA-rich diet produced divergent tissue-specific effects on SREBP-1c expression,<br />

significantly reducing hepatic SREBP-1c and increasing adipose tissue<br />

SREBP-1c expression, both of which could contribute to improved lipid and<br />

glucose metabolism. 28 Interestingly, this study also showed that TNF-α regulated<br />

SREBP-1c expression in human adipocytes, but not in hepatocytes, thus supporting<br />

the hypotheses that crosstalk exists between molecular markers of insulin<br />

sensitivity and adipocytokines which in turn can be modified by fatty acids.<br />

Further work showed that the insulin sensitizing effect of the c9, t11-CLArich<br />

diet was associated with a marked reduction in adipose tissue TNF-α expression<br />

that may be related to lower NF-κB DNA binding, which has been attributed<br />

to lower nuclear P65 levels and increased cytosolic inhibitor of κBα (IκBα)


246 Oxidative Stress and Inflammatory Mechanisms<br />

expression. 29 This study suggests that the fatty acid composition of the diet can<br />

be adjusted to attenuate the pro-inflammatory insulin de-sensitizing effect of<br />

obesity-induced insulin resistance. Indeed there was a significant reduction in the<br />

adipose tissue macrophage population observed in the c9, t11-CLA fed mice.<br />

This is an extremely important observation, in that it shows that dietary modification<br />

can alter the cellular profile of adipose tissue in obesity, which was<br />

associated with positive metabolic effects.<br />

WHOLE BODY METABOLIC PERSPECTIVE: EVIDENCE FROM<br />

HUMAN DIETARY FATTY ACID INTERVENTION STUDIES<br />

Several studies have shown a consistent relationship between plasma fatty acid<br />

composition and insulin resistance. A prospective cohort study investigated the<br />

interaction between serum fatty acid composition and the development of T2DM<br />

in a cohort of middle-aged normoglycemic men. 30 Baseline serum esterified and<br />

non-esterified SFA levels were significantly higher and PUFA levels were lower<br />

in the men who developed T2DM after 4 years.<br />

Recent evidence from the Nurses’ Health Study showed that higher intake of<br />

saturated fat and a low dietary P:S ratio were related to increased CVD risk<br />

among women with T2DM. 31 This study estimated that replacement of 5% of<br />

energy from saturated fat with equivalent energy from carbohydrates or MUFAs<br />

was associated with 22 and 37% lower risks of CVD, respectively. This finding<br />

suggests that dietary fatty acid modification may also determine secondary outcomes<br />

associated with metabolic syndrome and T2DM.<br />

Relatively few human dietary intervention studies have determined the relationship<br />

between dietary fatty acid composition and insulin sensitivity as the<br />

primary metabolic end point. For the purpose of this review, studies that investigated<br />

the effect of isocaloric substitution of dietary fatty acids will be reviewed,<br />

to exclude confounding effects of reduced energy intake on body weight. The<br />

KANWU study, a controlled multicenter, isoenergetic dietary intervention study<br />

involving 162 individuals, showed that decreasing dietary SFA and increasing<br />

MUFA improved insulin sensitivity but had no effect on insulin secretion. 32<br />

Interestingly the favorable effect of substituting SFA for MUFA was only seen<br />

when total fat intake was below 37% energy. Within each dietary group a second<br />

assignment of n-3 PUFA supplementation or placebo was completed, but the n-<br />

3 PUFA intervention had no effect on insulin sensitivity despite reduced TAG<br />

concentrations.<br />

A smaller randomized crossover dietary intervention study in 59 young<br />

healthy subjects randomly assigned to isoenergetic carbohydrate- and MUFArich<br />

diets for 28 days significantly improved insulin sensitivity compared to a<br />

high-SFA diet. 33 Also ex vivo analysis showed that both the carbohydrate- and<br />

MUFA-rich diets significantly increased basal and insulin-stimulated glucose<br />

uptake in monocytes. In contrast, another randomized, double-blind, crossover<br />

study comparing the effect of MUFA, SFA, and trans-fatty acid diets failed to


Dietary Fatty Acids and Metabolic Syndrome 247<br />

show any significant effects on insulin sensitivity or secretion. 34 When the group<br />

members were subdivided according to body mass index (BMI), insulin sensitivity<br />

was 24% lower in overweight individuals (BMI >25 kg/m 2 ) after the SFA diet<br />

compared to the MUFA diet. It is interesting to note that these diets were very<br />

low in fat (28% energy), Therefore the effects of dietary fat composition may be<br />

more obvious without the background low fat diet.<br />

Overall, human dietary intervention studies suggest that the removal of dietary<br />

saturated fat, as verified by alterations in plasma fatty acid composition, can have<br />

a direct effect on insulin sensitivity. This effect has also been confirmed in a<br />

metabolic study that showed that altering the compositions of infused free fatty<br />

acids affected insulin sensitivity. A SFA-rich lipid infusion significantly reduced<br />

insulin sensitivity indices (40 to 50%) to a much greater extent than a PUFA-rich<br />

lipid infusion (20 to 27%) in healthy subjects. 35<br />

Recent studies have attempted to determine whether alterations in dietary fat<br />

intake or endogenous fatty acid synthesis accounted for altered fatty acid composition<br />

associated with metabolic syndrome. Cross-sectional data suggest that<br />

insulin-resistant states are associated with high levels of activity of stearoyl-CoA<br />

desaturates (SCD-1) and Δ6-desaturase (D6D) and low Δ5-desaturase (D5D)<br />

activity. 36<br />

In a prospective study, Warensjo et al. 37 evaluated serum cholesteryl ester<br />

fatty acid composition and estimated SCD-1, D6D, and D5D activities as precursors<br />

to fatty acid ratios. The study showed that baseline fatty acid profiles<br />

predicted the development of metabolic syndrome 20 years later, SFA levels were<br />

significantly higher and linoleic acid levels were lower in subjects who subsequently<br />

developed metabolic syndrome by age 70. In addition SCD-1 and D6D<br />

activities were significantly higher and D5D activity was lower in those who<br />

developed metabolic syndrome during follow-up. The clinical relevance of altered<br />

desaturase activity in the development of the metabolic syndrome requires further<br />

study.<br />

Long chain n-3 PUFAs have a number of positive health benefits relevant to<br />

metabolic syndrome, particularly with respect to TAG metabolism. 36 However,<br />

we have relatively little evidence that n-3 PUFA supplementation improves insulin<br />

sensitivity in humans despite several studies showing that feeding n-3 PUFA had<br />

positive effects on glucose and insulin metabolism in different animal models of<br />

T2DM and metabolic syndrome. 39,40<br />

Also human epidemiological studies suggest that habitual dietary fish intake<br />

is inversely associated with the incidence of impaired glucose tolerance and<br />

T2DM. 41,42 Some studies have reported positive effects of n-3 PUFA supplementation<br />

on insulin sensitivity in individuals with impaired glucose tolerance and<br />

diabetes. 43,44 Other studies have not shown positive effects 32,45 even though n-3<br />

PUFA supplementation improved TAG metabolism. Clearly the putative effects<br />

of n-3 PUFA on human insulin resistance and impaired glucose tolerance require<br />

further clarification.


248 Oxidative Stress and Inflammatory Mechanisms<br />

FUTURE PERSPECTIVES<br />

The increasing prevalence of obesity requires us to reduce the impact of the<br />

adverse health effects, particularly T2DM and CVD. By 2010, some 31 million<br />

people in Europe and an estimated 239 million worldwide will require treatment<br />

for T2DM and its related complications. 46,47 The incidence of the metabolic<br />

syndrome is increasing exponentially as a consequence of the sharp rise in obesity<br />

— the key etiological factor in the development and severity of metabolic syndrome.<br />

To date, public health strategies have been largely unsuccessful at reducing<br />

the prevalence of obesity. Therefore dietary interventions that attenuate the severity<br />

of metabolic syndrome within the context of obesity are required and attenuating<br />

the impact of environmental causes through dietary fatty acid modification<br />

to improve insulin sensitivity would be advantageous.<br />

REFERENCES<br />

1. Alberti, K., Zimmet, P., and Consultation, W., Definition, diagnosis and classification<br />

of dibetes mellitus and its complications. Part 1: diagnosis and classification<br />

of diabetes mellitus, provisional report of a WHO consultation. Diabetic Med. 15,<br />

539, 1998.<br />

2. Executive Summary, Third Report of the National Cholesterol Education Programme<br />

(NCEP) Expert Panel on Detection, Evaluation and Treatment of High<br />

Blood Cholesterol in Adults (Adult Treatment Panel III). JAMA 285, 2486, 2001.<br />

3. Roche, H.M., Fatty acids and the metabolic syndrome. Proc. Nutr. Soc. 64, 23,<br />

2005.<br />

4. Magliano, J.D., Shaw, J.E., and Zimmet, P.Z., How best to define the metabolic<br />

syndrome. Ann. Med. 58, 34, 2006.<br />

5. Phinney, S.D., Fatty acids, inflammation and the metabolic syndrome. Am. J. Clin.<br />

Nutr. 82, 1151, 2005.<br />

6. Phillips, C.P. et al., Genetic and nutrient determinants of the metabolic syndrome,<br />

Curr. Op. Cardiol. 21, 185, 2006.<br />

7. Kahn, B.B. and Flier, J.S., Obesity and insulin resistance. J. Clin. Invest. 106,<br />

473, 2000.<br />

8. Roche, H.M., Phillips, C., and Gibney, M.J., The metabolic syndrome: the crossroads<br />

of diet and genetics. Proc. Nutr. Soc. 64, 371, 2005.<br />

9. Saltiel, A.R., The molecular and physiological basis of insulin resistance: emerging<br />

implications for metabolic and cardiovascular diseases. J. Clin. Invest. 106, 163,<br />

2000.<br />

10. Ghamin, H., Circulating mononuclear cells in the obese are in a proinflammatory<br />

state. Circulation 110, 1564, 2004.<br />

11. Dandona, P., Ajada, A., and Bandyopadhyay, A., Inflammation: the link between<br />

insulin resistance, obesity and diabetes. Trends Immunol. 25, 4, 2004.<br />

12. Roche, H.M., Dietary lipids and gene expression. Biochem. Soc. Trans. 32, 999,<br />

2004.<br />

13. Weisberg, S.P. et al., Obesity is associated with macrophage accumulation in<br />

adipose tissue. J. Clin. Invest. 112, 1796, 2004.


Dietary Fatty Acids and Metabolic Syndrome 249<br />

14. Xu, H. et al., Chronic inflammation in fat plays a crucial role in the development<br />

of obesity-related insulin resistance. J. Clin. Invest. 112, 1821, 2004.<br />

15. Griffin, M.E. et al., Free fatty acid-induced insulin resistance is associated with<br />

activation of protein kinase C-theta and alterations in the insulin signaling cascade.<br />

Diabetes 48, 1270, 1999.<br />

16. Shulman, G.I., Cellular mechanisms of insulin resistance. J. Clin. Invest. 106, 171,<br />

2000.<br />

17. Le Marchand-Brustel, Y. et al., Fatty acid-induced insulin resistance: role of insulin<br />

receptor substrate 1 serine phosphorylation in the retroregulation of insulin signalling.<br />

Biochem. Soc. Trans. 31, 1152, 2003.<br />

18. Wellen, K.E. and Hotamisigil, G.S., Inflammation, stress and diabetes. J. Clin.<br />

Invest. 115, 1111, 2005.<br />

19. Senn, J.J., et al., Suppressor of cytokine signaling-3 (SOCS-3), a potential mediator<br />

of interleukin-6-dependent insulin resistance in hepatocytes. J. Biol. Chem.<br />

278, 13740, 2003.<br />

20. Steppan, C.M. et al., The hormone resistin links obesity to diabetes. Nature 409,<br />

307, 2001.<br />

21. Hotamisligil, G.S., Inflammatory pathways and insulin action. Int. J. Obesity Rel.<br />

Metab Disord. 27, S53, 2003.<br />

22. Peraldi, P. and Spiegelman, B., TNF-alpha and insulin resistance: summary and<br />

future prospects. Mol. Cell. Biochem. 182, 169, 1998.<br />

23. Uysal, K.T., Weisbrock, S.M., and Hotamisligil, G.S., Functional analysis of<br />

tumor necrosis factor (TNF) receptors in TNF-alpha-mediated insulin resistance<br />

in genetic obesity. Endrocrinology 139, 4832, 1998.<br />

24. Kim, K.H. et al., A cysteine-rich adipose tissue-specific secretory factor inhibits<br />

adipocyte differentiation. J. Biol. Chem. 276, 11252, 2001.<br />

25. Hirosumi, J. et al., A central role for JNK in obesity and insulin resistance. Nature<br />

420, 333, 2002.<br />

26. Ajuwon, K.M. and Spurlock, M.E., Palmitate activates the NF-κB transcription<br />

factor and induces IL-6 and TNFα expression in 3T3-L1 adipocytes. J. Nutr. 135,<br />

1841, 2005.<br />

27. Nguyen, M.T. et al., JNK and tumor necrosis factor-α mediate free fatty acidinduced<br />

insulin resistance in 3T3-L1 adipocytes. J. Biol. Chem. 280, 35361, 2005.<br />

28. Roche, H.M. et al., Isomer-dependent metabolic effects of conjugated linoleic acid<br />

(CLA): insights from molecular markers SREBP-1c and LXRα. Diabetes 51,<br />

2037, 2002.<br />

29. Moloney, F. et al., Anti-diabetic effects of cis-9, trans-11 conjugated linoleic acid<br />

may be mediated via anti-inflammatory effects in white adipose tissue. Diabetes<br />

56(3), 574, 2007.<br />

30. Laaksonen, D.E. et al., Serum fatty acid composition predicts development of<br />

impaired fasting glycaemia and diabetes in middle-aged men. Diabetes Med. 19,<br />

456, 2002.<br />

31. Tanasescu, M. et al., Dietary fat and cholesterol and the risk of cardiovascular<br />

disease among women with type 2 diabetes. Am. J. Clin. Nutr. 79, 99, 2004.<br />

32. Vessby, B. et al., Substituting dietary saturated for monounsaturated fat impairs<br />

insulin sensitivity in healthy men and women: the KANWU study. Diabetologia<br />

44, 312, 2001.<br />

33. Perez-Jimenez, E. et al., A Mediterranean and a high-carbohydrate diet improve<br />

glucose metabolism in healthy young persons. Diabetologia, 44, 2038, 2001.


250 Oxidative Stress and Inflammatory Mechanisms<br />

34. Lovejoy, J.C. et al., Effects of diets enriched in saturated (palmitic), monounsaturated<br />

(oleic), or trans (eladic) fatty acids in insulin sensitivity and substrate<br />

oxidation in healthy adults. Diabetes Care 25, 1283, 2002.<br />

35. Stefan, N. et al., Effect of the pattern of elevated free fatty acids on insulin<br />

sensitivity and insulin secretion in healthy humans. Hormone Metab. Res. 33, 432,<br />

2001.<br />

36. Vessby, B. et al., Desaturation and elongation of fatty acids and insulin action.<br />

Ann. NY Acad. Sci. 967, 183, 2002.<br />

37. Warensjo, E., Riserus, U., and Vessby, B., Fatty acid composition of serum lipids<br />

predicts the development of the metabolic syndrome in men. Diabetologia, 48,<br />

1999, 2005.<br />

38. Roche, H.M. and Gibney, M.J., The effect of long-chain n-3 PUFA on fasting and<br />

postprandial triacylglycerol metabolism. Am. J. Clin. Nutr. 71, 232, 2000.<br />

39. Storlien, L.H. et al., Fish oil prevents insulin resistance induced by high fat feeding<br />

in rats. Science 237, 885, 1987.<br />

40. Aguilera, A.A. et al., Effects of fish oil on hypertension, plasma lipids and tumor<br />

necrosis factor-a in rats with sucrose-induced metabolic syndrome. J. Nutr. Biochem.<br />

15, 350, 2004.<br />

41. Feskens, E.J. et al., Dietary factors determining diabetes and impaired glucose<br />

tolerance: a 20-year follow-up of the Finnish and Dutch cohorts of the Seven<br />

Countries Study. Diabetes Care 18, 1104, 1985.<br />

42. Feskens, E.J., Bowles, C.H., and Kromhout, D., Inverse association between fish<br />

intake and risk of glucose intolerance in normoglycemic elderly men and women.<br />

Diabetes Care, 14, 935, 1991.<br />

43. Fasching, P. et al., Metabolic effects of fish-oil supplementation in patients with<br />

impaired glucose tolerance. Diabetes, 40, 583, 1991.<br />

44. Popp-Snijders, C. et al., Dietary supplementation of omega-3 polyunsaturated fatty<br />

acids improves insulin sensitivity in non-insulin-dependent diabetes. Diabetes Res.<br />

4, 141, 1987.<br />

45. Brady, L.M. et al., Increased n-6 polyunsaturated fatty acids do not attenuate the<br />

effects of long-chain n-3 polyunsaturated fatty acids on insulin sensitivity or<br />

triacylglycerol reduction in Indian Asians. Am. J. Clin. Nutr. 79, 983, 2004.<br />

46. King, H., Aubert, R.E., and Herman, W.H., Global burden of diabetes 1995–2025.<br />

Diabetes Care 21, 1414, 1998.<br />

47. Zimmet, P., Alberti, K.G.M.M., and Shaw, J., Global and societal implications of<br />

the diabetes epidemic. Nature 414, 782, 2001.


16<br />

CONTENTS<br />

Lipid-Induced Death of<br />

Macrophages:<br />

Implication for<br />

Destabilization of<br />

Atherosclerotic Plaques<br />

Oren Tirosh and Anna Aronis<br />

Abstract .............................................................................................................251<br />

Introduction .......................................................................................................252<br />

Macrophage Cell Death ....................................................................................253<br />

Programmed Cell Death..........................................................................253<br />

Cell Death and Atherosclerosis...............................................................254<br />

Effects of TGs on Macrophage Cell Deaths...........................................255<br />

Potential Therapies............................................................................................258<br />

References .........................................................................................................258<br />

ABSTRACT<br />

In recent decades, the incidence of atherosclerosis in Western society has been<br />

on the rise. Lipid particles, of which low density lipoprotein (LDL) is the most<br />

studied, accumulate in the intima of blood vessels and lead to inflammatory<br />

processes. Leukocytes, mostly macrophages, are recruited from circulating blood<br />

to the vessel walls. After they are overexposed to lipids and transform to foam<br />

cells, cell death processes take place. In macrophages, the exposure to lipids is<br />

known to trigger cell deaths of both apoptotic and necrotic types. External,<br />

receptor-dependent, and internal mitochondria-affecting apoptotic death pathways<br />

are involved in these processes. Although oxidized LDL is known as a<br />

major factor for plaque formation, triacylglycerols (TGs) are increasingly<br />

believed to be independently associated with coronary heart disease. Accumulation<br />

of intracellular TGs causes elevation of reactive oxygen species (ROS),<br />

251


252 Oxidative Stress and Inflammatory Mechanisms<br />

probably of mitochondrial sources, and cell death. In this chapter, the role of<br />

macrophage cell death in atherogenic plaque instability and the specific effects<br />

of TG in macrophage lipotoxicity are discussed.<br />

INTRODUCTION<br />

Coronary heart disease (CHD) is the largest cause of morbidity and mortality in<br />

the Western world. 1 Three cellular components of the circulation, monocytes,<br />

platelets, and T lymphocytes, together with two cell types of the artery wall cells,<br />

endothelial and smooth muscle cells (SMC), interact in multiple ways in concert<br />

with lipoprotein particles in generating atherosclerotic lesions. Accumulation and<br />

oxidation of LDL in vessel walls promotes up-regulation of adhesion molecules<br />

in endothelial cells and leads to recruitment of blood monocytes and lymphocytes<br />

to the intima. 2<br />

As shown by epidemiological and clinical studies, the Western way of life<br />

and consumption of an unbalanced diet lead to a high incidence of atherosclerosis<br />

and conditions significantly increasing the risk of CHD such as obesity, metabolic<br />

syndrome, and diabetes. 3 An atherogenic shift in blood lipid spectrum profile that<br />

may manifest by increases of serum levels of cholesterol, LDLs, TGs, and very<br />

low density lipoproteins (VLDLs), and decreases of serum levels of anti-atherogenic<br />

HDLs has become a frequent problem in developed countries.<br />

Recruitment of immune cells to the intima of blood vessels indicates an<br />

inflammation reaction caused by lipid particles. Indeed, obese persons have<br />

increased serum or plasma concentrations of acute-phase proteins or pro-inflammatory<br />

cytokines such as C-reactive protein (CRP), interleukin (IL)-6, IL-8, and<br />

tumor necrosis factor (TNF). Circulating levels of these immune mediators can<br />

be lowered by weight loss 4,5 or other plasma lipid-lowering interventions, for<br />

example, short-term diet and exercise. 6,7 CRP, which has been reported to be<br />

expressed by the liver, macrophages and SMC-like cells, correlates with macrophage<br />

accumulation in coronary arteries. 8 Therefore, a reduction of CRP by lipidlowering<br />

measures can contribute to prevention of macrophage accumulation in<br />

atherosclerotic plaques.<br />

LDL cholesterol is currently defined as a major risk factor of CHD. However,<br />

the role of TG, which is increasingly believed to be independently associated<br />

with CHD, is not yet well studied. A high blood level of TG often occurs together<br />

with low HDL. Such an effect on the blood lipid profile often occurs with normal<br />

levels of LDL-C. These abnormalities of the TG-HDL axis are characteristic of<br />

patients with metabolic syndrome. 9 Recent clinical studies have also revealed that<br />

increased serum triglyceride (TG) levels are closely related to atherosclerosis<br />

independently of serum levels of HDL and LDL.<br />

Among TG-rich lipoproteins (TRLs), remnant lipoproteins (RLPs) are considered<br />

atherogenic and independent coronary risk factors. It was previously<br />

reported 10 that monocytes cultured in the presence of RLPs increased their adhesion<br />

capability to vascular endothelial cells. In the Third Report of the Adult<br />

Treatment Panel (ATP III) establishing the National Cholesterol Education


Lipid-Induced Death of Macrophages 253<br />

Program (NCEP) in the United States, the attitude toward elevated plasma triglyceride<br />

levels has changed, especially as related to moderately elevated and<br />

borderline high levels. TGs were cited as independent risk factors for CHD.<br />

PROGRAMMED CELL DEATH<br />

MACROPHAGE CELL DEATH<br />

Apoptosis is often associated with morphological and biochemical changes. 11,12<br />

During apoptosis, the nucleus and cytoplasm are condensed and the dying cells<br />

disintegrate into membrane-bound apoptotic bodies. Nucleases are activated and<br />

cause the degradation of chromosomal DNA, at first into large chromosomal<br />

DNA (50 to 300 kb) and ultimately into very small oligonucleosomal fragments.<br />

11,12 The signaling events leading to apoptosis can be divided into two<br />

distinct pathways involving either mitochondria or death receptors. 13,14<br />

In the mitochondrial pathway, death signals led to changes in mitochondrial<br />

membrane permeability and the subsequent release of pro-apoptotic factors<br />

involved in various aspects of apoptosis. 13 The released factors included cytochrome<br />

c (cyt c), 15 apoptosis inducing factor (AIF), 16 second mitochondriaderived<br />

activator of caspase (Smac/DIABLO), 17,18 and endonuclease G. 19 Cytosolic<br />

cyt c forms an essential part of the apoptosome apoptosis complex composed<br />

of cyt c, Apaf-1, and procaspase-9. Formation of the apoptosome leads to the<br />

activation of caspase-9, which then processes and activates other caspases to<br />

orchestrate the biochemical executions of cells. Smac/DIABLO is also released<br />

from the mitochondria along with cyt c during apoptosis, and it functions to<br />

promote activation of caspases by inhibiting IAP (inhibitor of apoptosis) family<br />

proteins. 17,18<br />

In the death receptor pathway, the apoptotic events are initiated by engaging<br />

the tumor necrosis factor (TNF) family receptors including eight different death<br />

domain (DD)-containing receptors (TNFR1 also called DR1; Fas, also called<br />

DR2; DR3, DR4, DR5, DR6, NGFR, and EDAR). 20,21 Upon ligand binding or<br />

when overexpressed in cells, DD receptors aggregate, resulting in the recruitment<br />

of various adapter proteins that mediate both cell death and proliferation 20,21 and<br />

can activate the caspase system. The human genome encodes 12 to 13 distinct<br />

caspases that function in cytokine processing and inflammation, and at least seven<br />

(caspases 2, 3, 6, 7, 8, 9, and 10) contribute to cell death. 22 To date more than<br />

280 caspase targets have been identified. 23 Some of these proteins may be cleaved<br />

very late and less completely during apoptosis or may not be cleaved in all cell<br />

types. The functional consequences of the cleavage of many of the identified<br />

substrates are unknown. 23 Most of the protein substrates can be categorized into<br />

a few functional groups: apoptosis regulator, cell adhesion, cytoskeletal and<br />

structural, etc. Rucci et al. noted that proteins of the electron transfer chains of<br />

mitochondria are targets for caspase-dependent degradation during apoptosis. 24<br />

Active caspase-9 and caspase-3 have been observed in the mitochondria, but<br />

their origins are unclear. Theoretically, procaspase-9 may be activated in the


254 Oxidative Stress and Inflammatory Mechanisms<br />

mitochondria in a cytochrome c/Apaf-1-dependent manner, or activated caspase-<br />

9 and -3 may translocate to the mitochondria as suggested by Chandra and Tang. 25<br />

Using a system of positive staining of the mitochondria with calcein and cobalt<br />

as a fluorescent quencher, Poncet et al. showed that during induction of apoptosis,<br />

the inner membrane of the mitochondria losses its barrier function and becomes<br />

permeable. 26 These data suggest that caspase activity can translocate to the mitochondrial<br />

matrix. Indeed strong support for this idea is based on detection of<br />

mitochondrial caspase activity in real time, in situ, in live cells. Using a mitochondrially<br />

targeted CFP-caspase 3 substrate-YFP construct (mC3Y), a caspase-<br />

3-like activity in the mitochondrial matrices of some cells during apoptosis was<br />

demonstrated. 27<br />

CELL DEATH <strong>AND</strong> ATHEROSCLEROSIS<br />

One of the earliest events in atherosclerosis is the entry of monocytes into focal<br />

areas of the arterial subendothelium. Once inside the arterial intima, monocytes<br />

differentiate into macrophages. 28 Macrophages are constantly exposed to excess<br />

lipids following overnutrition and they are also involved in the development of<br />

atherosclerosis — an inflammatory disease process. 29–31 Macrophages are considered<br />

to constitute one of the most important factors in its initiation and progression.<br />

Macrophages accumulate large amounts of intracellular cholesterol via<br />

accumulation of lipoproteins. The presence of cholesterol-loaded macrophages<br />

in atherosclerotic lesions is a prominent feature throughout the lives of the lesions<br />

and these cells exert major impacts on lesion progression.<br />

Dead macrophages are frequently observed in human atherosclerotic lesions,<br />

and are considered to be involved in atherosclerotic plaque instability. 32 Unstable<br />

plaques demonstrate a greater portion of apoptotic cells than stable ones. 33,34 Immunohistochemical<br />

staining of ruptured plaques has shown that apoptotic nuclei in<br />

plaque rupture sites are essentially those of macrophages and much less frequently<br />

of smooth muscle and T lymphocytes. 33 Fibrous caps of ruptured plaques have more<br />

macrophages and also contain fewer SMCs than those of unruptured ones. The<br />

death of macrophages and SMCs in blood vessels is a complicated process owing<br />

to their mutual interactions and the presence of monocytes and macrophages in<br />

plaques increases also the rate of SMC apoptosis. 35<br />

Accumulated evidence suggests that plaque macrophage deaths are due to<br />

three possible pathways: (1) Fas-mediated death and promotion of inflammation,<br />

(2) oxidized LDL which at least in vitro was proven to be an effective cell death<br />

inducer following cellular uptake, and (3) excess accumulation of free cholesterol<br />

that may lead to caspase-dependent or independent cell death. The level of<br />

apoptotic cell death is strongly related to the stage of development of the atherosclerotic<br />

plaque. 36 It has been suggested that in the early stages of lesions, the<br />

induction of macrophages apoptosis actually prevents the growth and development<br />

of the lesion. This effect is probably attributed to controlling the number<br />

of macrophages in the lesion. The experimental evidence of this effect of apoptosis<br />

includes a number of genetic alterations in mouse models of atherosclerosis


Lipid-Induced Death of Macrophages 255<br />

that resulted in early increases or decreases of macrophage apoptosis. An inverse<br />

relationship was found between early lesional macrophage apoptosis and lesion<br />

area.<br />

When bone marrow taken from P53–/– (a pro-apoptotic protein) mice was<br />

transplanted to APOE*3-Leiden mice, a significant 2.3-fold increase was<br />

observed in early lesion size. When bone marrow of Bax–/– mice was transplanted<br />

to LDLR–/– mice, the same effect was observed, indicating a negative<br />

effect of apoptosis on plaque growth. On the other hand, in an advanced atherosclerotic<br />

plaque region, the apoptotic processes may lead to massive macrophage<br />

cell deaths. Impaired capacities of other macrophages to clean and clear<br />

the plaque regions from cell debris may lead to the development of necrotic<br />

cores and accelerated inflammatory processes in the intima. Macrophages<br />

express multiple metalloproteinases (e.g., stromelysin) and serine proteases<br />

(e.g., urokinase) that degrade the extracellular matrix, weakening the plaque and<br />

making it rupture-prone.<br />

Most studies report apoptotic macrophage deaths in ruptured plaques on the<br />

basis of immunohistochemical staining methods. However, standard TUNEL<br />

DNA staining lacks specificity, especially in late phases of cell death when most<br />

DNA passes degradation. 36 Moreover, because of variable uptake of lipids and<br />

differences in cell size, apoptosis-characterizing cell shrinkage is not always<br />

apparent. 33 This is why an indication of the active form of “committing-to-die”<br />

caspase-3 is necessary for final determination of apoptotic cell death. Some<br />

researchers indicate caspase-3 independent cell deaths in LDL-exposed macrophage<br />

cell cultures or histological cuts of plaques, 33,37 supposing an alternative<br />

of oncotic (necrotic) cell death.<br />

EFFECTS OF TGS ON MACROPHAGE CELL DEATHS<br />

Previous studies have suggested that TGs are the main signaling components of<br />

endocytosed VLDL in macrophages. 38 Our study showed death pathways in<br />

macrophages resulting from exposure to TGs — a mechanism that may be relevant<br />

to the development of atherosclerosis. Murine J774.2 macrophages in culture<br />

were exposed to a commercial soybean oil-based TG lipid emulsion (0.1 to 1.5<br />

mg lipid/ml) known to be taken up by macrophages via a coated pit-dependent<br />

mechanism mediated by macrophage secretion of apolipoprotein E. 40 The exposure<br />

of the cells to the lipid emulsion led to intracellular change in fatty acid<br />

profile and high accumulation of TGs as measured with gas chromatography and<br />

Nile red staining, respectively (Figure 16.1). The TG effect culminated in cell<br />

death, with no caspase-3 activation. Dual staining with propidium iodide and<br />

annexin V followed by flow cytometric analysis showed that TG facilitated cell<br />

deaths with clear necrotic characteristics.<br />

Reactive oxygen species (ROS) are involved in macrophage activation and<br />

death. While ROS may induce macrophage activation, macrophage foam cells<br />

contain potent oxidant-generating lipid-targeting systems such as inducible NO


256 Oxidative Stress and Inflammatory Mechanisms<br />

Events<br />

128<br />

C<br />

12h<br />

0<br />

10<br />

Nile Red fluorescence<br />

0 101 102 103 104 FIGURE 16.1 Accumulation of intracellular lipids following exposure of J774.2 macrophages<br />

to lipid treatment. Nile red staining of the cells was carried out after treatment<br />

with soybean oil-based lipid emulsion, 1 mg/ml.<br />

synthase and 15-lipoxygenase, allowing increased recognition and uptake by<br />

macrophages and creating a positive feedback loop. 41<br />

Following 24 hr of exposure of macrophages to 1 mg/ml TGs, cellular ROS<br />

levels were strongly elevated. In contrast, after 48 hr, when 50% of the macrophages<br />

underwent cell death, ROS production was arrested. Most of the TGmediated<br />

ROS production was demonstrated to be via mitochondrial complex 1<br />

of the electron transfer chain, as demonstrated by the use of rotenone, a mitochondrial<br />

complex 1 inhibitor that significantly attenuated cellular ROS levels in<br />

TG-treated cells (Scheme 16.1).<br />

To elucidate whether the cell death process was indeed oxidant dependent,<br />

antioxidant protection was studied. Treatment with 0.5 mM N-acetyl-cysteine<br />

(NAC), 0.05 mM ascorbic acid, and 0.2 mM resveratrol protected against the TG<br />

lipotoxic effect, while lipophilic antioxidants surprisingly did not. For further<br />

study, the combination of NAC, ascorbic acid, and resveratrol was used at much<br />

lower concentrations (one-tenth of original concentrations), which led to the<br />

appearance of a synergistic protective effect.<br />

Exposure of J774.2 macrophages to increased levels of TG leads to the<br />

induction of oxidative stress-mediated lipotoxicity. Decreased GSH levels and<br />

the protective role of NAC are strong indications of the pivotal role of ROS in<br />

TG-induced lipotoxicity. In our model, the NAC, ascorbic acid, and resveratrol<br />

antioxidants played a protective role in TG-induced lipotoxicity. The protective<br />

effect of NAC, which is a precursor of glutathione, suggests an interaction in the<br />

level of water-soluble antioxidants. Moreover, TG caused ROS-dependent G1<br />

arrest in the macrophage cell cycle. The exposure of untreated cells to an inhibitor<br />

of glutathione synthesis BSO mimicked TG-induced G1 arrest, reinforcing an<br />

important role of GSH in prevention of TG-induced mechanisms of lipotoxicity.<br />

24h<br />

48h


Lipid-Induced Death of Macrophages 257<br />

Extra cellular<br />

Intra cellular<br />

TG particle<br />

ROS<br />

Caspase-3<br />

GSH<br />

Cell death Necrosis<br />

SCHEME 16.1 Mechanism for TG-induced cell death in macrophages. Lipid droplets<br />

accumulate inside the macrophages, in time generating foam cells. Oxidative stress develops<br />

as a result of the lipid stress and leads to a necrotic type of cell death. Overall, the<br />

effect of macrophage disintegration in atherogenic plaques will result in inflammation and<br />

plaque rupture due to destabilization.<br />

It is possible that prolonging ROS production from endogenous cellular<br />

sources such as mitochondria may attenuate caspase activity and shift apoptosis<br />

to necrosis. We therefore evaluated the effect of TG on apoptotic cells showing<br />

high caspase activity. Indeed, TG induced elevated ROS levels and suppressed<br />

caspase-3 in apoptotic cells pretreated for 24 hr with cycloheximide. These results<br />

indicate that exposure to TG can directly regulate lipotoxicity in macrophages<br />

by inducing mitochondria-mediated prolonged oxidative stress; this, in turn, can<br />

inactivate the apoptotic caspase system, resulting in necrotic cell death which can<br />

be prevented by specific antioxidants.<br />

Accumulation of TG in macrophages is a unique phenomenon of these cells<br />

since they are capable of taking up large amounts of this type of lipid. We<br />

suggest that rapid accumulation of fat may result in oxidative stress-dependent<br />

suppression of the caspase system which may suppress activation of type 1<br />

programmed cell death and affect other mechanisms related to cellular signal<br />

transduction and cell cycle arrest. The TG treatment may promote characteristics<br />

of necrotic cell death or caspase-independent types of programmed cell<br />

death. This, in fact, may result in disintegration of lipid-loaded macrophages<br />

(foam cells) and release of their enzymatic cytosolic and lysosomal contents<br />

to their surroundings, leading to lesion erosion and rupture of the plaques with<br />

the formation of a thrombus.<br />

?<br />

Mitochondria<br />

Inflammation


258 Oxidative Stress and Inflammatory Mechanisms<br />

POTENTIAL THERAPIES<br />

One problem that must be encountered when considering novel therapies in<br />

atherosclerosis is the heterogeneity of plaque status. In human patients, it is most<br />

probable that advanced plaques and early plaques are relatively abundant. Therefore,<br />

a careful approach should be considered. It is recommended that apoptosis<br />

will be promoted in the early plaques while in advanced plaques where inflammation<br />

is already a considerable factor, apoptosis should be inhibited. 42<br />

One approach for blocking macrophage cell death is to use specific antioxidants.<br />

We found that water-soluble antioxidants are more efficient than lipidsoluble<br />

antioxidants in preventing TG-induced lipotoxicity in macrophages.<br />

Therefore the use of antioxidants should be considered according to their relative<br />

death-preventing effects. Only compounds that may affect lipotoxicity should be<br />

used. Another approach is to enhance the phagocytic activities of macrophages<br />

in atherosclerotic plaques. This approach will help to maintain phagocytic activities<br />

in advance plaques and will prevent the formation of necrotic cores and the<br />

inflammatory responses. Use of eicosanoids including LXA4, LXB4, and aspirintriggered<br />

15-epi-LXB4, and their stable analogues stimulated macrophages to<br />

phagocytose apoptotic neutrophils. Apoplipoprotein E3 can correct defects in<br />

macrophage phagocytic effects and finally the use of glucocorticoids may boost<br />

phagocytic activity. 42<br />

REFERENCES<br />

1. Shah, P.K. (2003) Pathophysiology of plaque rupture and the concept of plaque<br />

stabilization, Cardiol. Clin. 21, 303.<br />

2. Osterud, B. and Bjorklid, E. (2003) Role of monocytes in atherogenesis, Physiol.<br />

Rev. 83, 1069.<br />

3. Gasbarrini, A. and Piscaglia, A.C. (2005) A natural diet versus modern Western<br />

diets? A new approach to prevent “well-being syndromes,” Dig. Dis. Sci. 50, 1.<br />

4. Herder, C. et al. (2006) Systemic monocyte chemoattractant protein-1 concentrations<br />

are independent of type 2 diabetes or parameters of obesity: results from<br />

the Cooperative Health Research in the Region of Augsburg Survey S4 (KORA<br />

S4), Eur. J. Endocrinol. 154, 311.<br />

5. Heilbronn, L.K. and Clifton, P.M. (2002) C-reactive protein and coronary artery<br />

disease: influence of obesity, caloric restriction and weight loss, J. Nutr. Biochem.<br />

13, 316.<br />

6. Roberts, C.K. et al. (2006) Effect of a short-term diet and exercise intervention<br />

on oxidative stress, inflammation, MMP-9, and monocyte chemotactic activity in<br />

men with metabolic syndrome factors, J. Appl. Physiol. 100, 1657.<br />

7. Kasim-Karakas, S.E., Tsodikov, A., Singh, U., and Jialal, I. (2006) Responses of<br />

inflammatory markers to a low-fat, high-carbohydrate diet: effects of energy<br />

intake, Am. J. Clin. Nutr. 83, 774.<br />

8. Turk, J.R. et al. (2003) C-reactive protein correlates with macrophage accumulation<br />

in coronary arteries of hypercholesterolemic pigs, J. Appl. Physiol. 95, 1301.


Lipid-Induced Death of Macrophages 259<br />

9. Szapary, P.O. and Rader, D.J. (2004) The triglyceride-high-density lipoprotein<br />

axis: an important target of therapy? Am. Heart J. 148, 211.<br />

10. Kawakami, A. and Yoshida, M. (2005) Remnant lipoproteins and atherogenesis,<br />

J. Atheroscler. Thromb. 12, 73.<br />

11. Steller, H. (1995) Mechanisms and genes of cellular suicide, Science 267, 1445.<br />

12. Nagata, S. (1997) Apoptosis by death factor, Cell 88, 355.<br />

13. Green, D.R. and Reed, J.C. (1998) Mitochondria and apoptosis, Science 281, 1309.<br />

14. Ashkenazi, A. and Dixit, V.M. (1998) Death receptors: signaling and modulation,<br />

Science 281, 1305.<br />

15. Liu, X., Kim, C.N., Yang, J., Jemmerson, R., and Wang, X. (1996) Induction of<br />

apoptotic program in cell-free extracts: requirement for dATP and cytochrome c,<br />

Cell 86, 147.<br />

16. Susin, S.A. et al. (1999) Molecular characterization of mitochondrial apoptosisinducing<br />

factor, Nature 397, 441.<br />

17. Du, C., Fang, M., Li, Y., Li, L., and Wang, X. (2000) Smac, a mitochondrial<br />

protein that promotes cytochrome c-dependent caspase activation by eliminating<br />

IAP inhibition, Cell 102, 33.<br />

18. Verhagen, A.M. et al. (2000) Identification of DIABLO, a mammalian protein that<br />

promotes apoptosis by binding to and antagonizing IAP proteins, Cell 102, 43.<br />

19. Li, L.Y., Luo, X., and Wang, X. (2001) Endonuclease G is an apoptotic DNase<br />

when released from mitochondria, Nature 412, 95.<br />

20. Bhardwaj, A. and Aggarwal, B.B. (2003) Receptor-mediated choreography of life<br />

and death, J. Clin. Immunol. 23, 317.<br />

21. Singh, A., Ni, J., and Aggarwal, B.B. (1998) Death domain receptors and their<br />

role in cell demise, J. Interferon Cytokine Res. 18, 439.<br />

22. Kroemer, G. and Martin, S.J. (2005) Caspase-independent cell death, Nat. Med.<br />

11, 725.<br />

23. Fischer, U., Janicke, R.U., and Schulze-Osthoff, K. (2003) Many cuts to ruin: a<br />

comprehensive update of caspase substrates, Cell Death Differ. 10, 76.<br />

24. Ricci, J.E. et al. (2004) Disruption of mitochondrial function during apoptosis is<br />

mediated by caspase cleavage of the p75 subunit of complex I of the electron<br />

transport chain, Cell 117, 773.<br />

25. Chandra, D. and Tang, D.G. (2003) Mitochondrially localized active caspase-9<br />

and caspase-3 result mostly from translocation from the cytosol and partly from<br />

caspase-mediated activation in the organelle: lack of evidence for Apaf-1-mediated<br />

procaspase-9 activation in the mitochondria, J. Biol. Chem. 278, 17408.<br />

26. Poncet, D., Boya, P., Metivier, D., Zamzami, N., and Kroemer, G. (2003) Cytofluorometric<br />

quantitation of apoptosis-driven inner mitochondrial membrane permeabilization,<br />

Apoptosis 8, 521.<br />

27. Zhang, Y. et al. (2004) Detection of mitochondrial caspase activity in real time in<br />

situ in live cells, Microsc. Microanal. 10, 442.<br />

28. Linton, M.F. and Fazio, S. (2003) Macrophages, inflammation, and atherosclerosis,<br />

Int. J. Obes. Relat. Metab. Disord. 27, Suppl. 3, S35.<br />

29. Hansson, G.K., Zhou, X., Tornquist, E., and Paulsson, G. (2000) The role of<br />

adaptive immunity in atherosclerosis, Ann. NY Acad. Sci. 902, 53.<br />

30. Patrick, L. and Uzick, M. (2001) Cardiovascular disease: C-reactive protein and<br />

the inflammatory disease paradigm: HMG-CoA reductase inhibitors, alpha-tocopherol,<br />

red yeast rice, and olive oil polyphenols: a review of the literature, Altern.<br />

Med. Rev. 6, 248.


260 Oxidative Stress and Inflammatory Mechanisms<br />

31. Isomaa, B. (2003) A major health hazard: the metabolic syndrome, Life Sci. 73,<br />

2395.<br />

32. Feng, B., Zhang, D., Kuriakose, G., Devlin, C.M., Kockx, M., and Tabas, I. (2003)<br />

Niemann-Pick C heterozygosity confers resistance to lesional necrosis and macrophage<br />

apoptosis in murine atherosclerosis, Proc. Natl. Acad. Sci. USA 100, 10423.<br />

33. Kolodgie, F.D. et al. (2000) Localization of apoptotic macrophages at the site of<br />

plaque rupture in sudden coronary death, Am. J. Pathol. 157, 1259.<br />

34. Moreno, P.R., Falk, E., Palacios, I.F., Newell, J.B., Fuster, V., and Fallon, J.T.<br />

(1994) Macrophage infiltration in acute coronary syndromes: implications for<br />

plaque rupture, Circulation 90, 775.<br />

35. Seshiah, P.N. et al. (2002) Activated monocytes induce smooth muscle cell death:<br />

role of macrophage colony-stimulating factor and cell contact, Circulation 105,<br />

174.<br />

36. Kockx, M.M. (1998) Apoptosis in the atherosclerotic plaque: quantitative and<br />

qualitative aspects, Arterioscler. Thromb. Vasc. Biol. 18, 1519.<br />

37. Baird, S.K., Reid, L., Hampton, M.B., and Gieseg, S.P. (2005) OxLDL-induced<br />

cell death is inhibited by the macrophage synthesised pterin, 7,8-dihydroneopterin,<br />

in U937 cells but not THP-1 cells, Biochim. Biophys. Acta 1745, 361.<br />

38. Chawla, A. et al. (2003) PPARdelta is a very low-density lipoprotein sensor in<br />

macrophages, Proc. Natl. Acad. Sci. USA 100, 1268.<br />

39. Aronis, A., Madar, Z., and Tirosh, O. (2005) Mechanism underlying oxidative<br />

stress-mediated lipotoxicity: exposure of J774.2 macrophages to triacylglycerols<br />

facilitates mitochondrial reactive oxygen species production and cellular necrosis,<br />

Free Radic. Biol. Med. 38, 1221.<br />

40. Carvalho, M.D. et al. (2002) Macrophages take up triacylglycerol-rich emulsions<br />

at a faster rate upon co-incubation with native and modified LDL: investigation<br />

on the role of natural chylomicrons in atherosclerosis, J. Cell. Biochem. 84, 309.<br />

41. Bennett, M.R. (2001) Reactive oxygen species and death: oxidative DNA damage<br />

in atherosclerosis, Circ. Res. 88, 648.<br />

42. Tabas, I. (2005) Consequences and therapeutic implications of macrophage apoptosis<br />

in atherosclerosis: the importance of lesion stage and phagocytic efficiency,<br />

Arterioscler. Thromb. Vasc. Biol. 25, 2255.


17<br />

CONTENTS<br />

α-Lipoic Acid Prevents<br />

Diabetes Mellitus and<br />

Endothelial Dysfunction<br />

in Diabetes-Prone<br />

Obese Rats<br />

Woo Je Lee, Ki-Up Lee, and Joong-Yeol Park<br />

Introduction .......................................................................................................261<br />

α-Lipoic Acid....................................................................................................262<br />

α-Lipoic Acid Prevents Diabetes Mellitus in Diabetes-Prone<br />

Obese Rats...................................................................................263<br />

α-Lipoic Acid Prevents Endothelial Dysfunction in Diabetes-Prone<br />

Obese Rats...................................................................................265<br />

Summary ...........................................................................................................267<br />

Acknowledgment...............................................................................................267<br />

References .........................................................................................................267<br />

INTRODUCTION<br />

Recent evidence suggests that oxidative stress plays a causative role in many<br />

chronic diseases such as diabetes and cardiovascular disease. Many efforts have<br />

been made to prevent the development and progression of diabetes and cardiovascular<br />

disease by using materials with antioxidative properties. Among a variety<br />

of antioxidants, α-lipoic acid (α-LA) has recently received much research attention.<br />

This chapter will focus on the roles and mechanisms of α-LA in the<br />

prevention of diabetes mellitus and endothelial dysfunction in obese animals.<br />

261


262 Oxidative Stress and Inflammatory Mechanisms<br />

α-LIPOIC ACID<br />

α-LA is a naturally occurring compound that is widely distributed in plants and<br />

animals. In humans, α-LA can be supplied by diet or obtained through de novo<br />

synthesis by the liver and other tissues. Only the R-isomer of α-LA is synthesized<br />

naturally. Conventional chemical synthesis of α-LA results in a 50/50 (racemic)<br />

mixture of two optical isomers, R-α-LA and S-α-LA.<br />

Free α-LA is rapidly taken up by cells and reduced to dihydrolipoic acid<br />

(DHLA) intracellularly. In most cells containing mitochondria, α-LA is reduced<br />

by an NADH-dependent reaction with lipoamide dehydrogenase to form DHLA.<br />

In cells that lack mitochondria, α-LA can be reduced to DHLA via NADPH with<br />

glutathione and thioredoxin reductase. 1<br />

α-LA contains two sulfur molecules that can be oxidized or reduced (Figure<br />

17.1). This feature allows α-LA to function as a cofactor for several important<br />

enzymes as well as a potent antioxidant. α-LA functions as a cofactor of mitochondrial<br />

key enzymes such as pyruvate dehydrogenase and α-keto-glutarate<br />

dehydrogenase and thus has been shown to be required for the oxidative decarboxylation<br />

of pyruvate to acetyl-CoA, the critical step leading to the production<br />

of cellular energy (ATP). 2,3 In addition, when large amounts of free α-LA are<br />

available (e.g., with supplementation), α-LA is also able to function as an antioxidant.<br />

4 α-LA and its reduced form, DHLA, are powerful antioxidants. 5 DHLA<br />

can then be transported easily out of the interiors of cells and function effectively<br />

in the extracellular spaces. Thus, in this form, it is believed to function directly<br />

as an antioxidant and possess its greatest antioxidant potential. 6 However, because<br />

DHLA is also rapidly eliminated from cells, the extents to which its antioxidant<br />

effects can be sustained remain unclear.<br />

Many reports describe the functions of α-LA 4,5,7–10 and the functions include<br />

(1) quenching of reactive oxygen species, (2) regeneration of exogenous and<br />

endogenous antioxidants such as vitamins C and E and glutathione, (3) chelation<br />

of metal ions, and (4) prevention of membrane lipid peroxidation and reparation<br />

of oxidized proteins. Because of its antioxidant properties, α-LA is known to be<br />

effective in both prevention and treatment of oxidative stress in a number of<br />

Lipoic acid<br />

Dihydrolipoic acid<br />

(DHLA)<br />

S — S<br />

FIGURE 17.1 Structures of lipoic acid and dihydrolipoic acid.<br />

∗<br />

SH SH<br />

∗<br />

O<br />

O<br />

OH<br />

OH


Diabetes Mellitus and Endothelial Dysfunction 263<br />

models or clinical conditions including diabetes, 11–14 vascular dysfunction, 15 and<br />

neurodegenerative diseases. 16<br />

α-LIPOIC ACID PREVENTS DIABETES MELLITUS IN DIABETES-PRONE<br />

OBESE RATS<br />

Increased oxidative stress is a widely accepted participant in the development<br />

and progression of diabetes and its complications. 17–19 Diabetes is usually accompanied<br />

by increased production of free radicals 18–21 or impaired antioxidant<br />

defenses. 22,23 Hence, it is likely that a substance known to reduce oxidative stress<br />

would reduce the progression of cell damage in diabetes.<br />

α-LA, an essential cofactor in oxidative metabolism, has been reported to<br />

exert a number of potentially beneficial effects in oxidative stress-related conditions.<br />

Experimental and clinical studies have indicated that α-LA treatment<br />

reduces the development of diabetic complications. 7,24–29 Furthermore, α-LA<br />

facilitates glucose metabolism and increases glucose uptake leading to improved<br />

glucose utilization in vitro and in vivo. 30 These features may enable α-LA to be<br />

used as a potential agent for the prevention and treatment of diabetes mellitus.<br />

Experimental and clinical studies have indicated that α-LA improved insulin<br />

sensitivity. α-LA administration improved insulin-stimulated whole body and<br />

skeletal muscle glucose utilization in insulin-resistant obese rats 30,31 and type 2<br />

diabetic humans. 32,33 Moreover, several studies using cultured muscle cells, 34<br />

isolated rat diaphragm, 35 and perfused preparations of normal and diabetic<br />

rabbits 36 demonstrated that α-LA activates glucose transport and stimulates<br />

glycolysis.<br />

Obesity is the most important risk factor for type 2 diabetes mellitus. 37 Many<br />

possible mechanisms link obesity and the development of diabetes mellitus. One<br />

mechanism is that oxygen free radicals derived from excessive triglycerides and<br />

long chain fatty acyl CoA (LCAC) in muscles and pancreatic β cells cause<br />

functional defects in these tissues, leading to the development of type 2 diabetes.<br />

38–40 α-LA has been shown to protect against oxidative stress-induced insulin<br />

resistance in vitro 41,42 and improve insulin sensitivity in human and animal models.<br />

30,43–47 In addition, it was recently reported that administration of α-LA to<br />

rodents reduced body weight by suppressing food intake and visceral fat mass. 48<br />

Thus, it is conceivable that α-LA may have a preventive role in the development<br />

of diabetes.<br />

A recent study (14) demonstrated that α-LA could prevent the development<br />

of diabetes in an obese animal model. The authors used Otsuka Long-Evans<br />

Tokushima Fatty (OLETF) rats as obese animal models, because these rats are<br />

obese from a young age and become diabetic after 18 weeks of age. 49 These<br />

features in OLETF rats resemble those of human type 2 diabetes. To investigate<br />

the preventive effect of α-LA on the development of diabetes, 9-week old OLETF<br />

rats were fed standard rat chow with or without racemic α-LA (200 mg/kg of<br />

body weight/day) for 3 weeks and the development of diabetes was evaluated.<br />

Approximately 80% of rats fed rat chow without α-LA developed diabetes at 40


264 Oxidative Stress and Inflammatory Mechanisms<br />

weeks of age, but none of the rats fed rat chow with α-LA developed diabetes.<br />

In addition, administration of α-LA protected pancreatic β cell destruction and<br />

reduced triglyceride accumulations in skeletal muscles and pancreatic β cells.<br />

These data indicate that α-LA prevents the development of diabetes in diabetesprone<br />

obese rats by decreasing lipid accumulation in skeletal muscle and exerting<br />

beneficial effects on pancreatic β cells.<br />

This report seems to indicate that the mechanism of preventive effect of α-<br />

LA on the development of diabetes may be multifarious. First, because oxidative<br />

stress has been suggested to be associated with insulin resistance 50,51 and α-LA<br />

has potent antioxidative capacities, its preventive effects may be due to its antioxidant<br />

action. Indeed, α-LA reduced plasma levels of oxidative stress markers<br />

such as malondialdehyde and 8-hydroxy-deoxyguanosine. However, this relationship<br />

between protective effect and antioxidant action of α-LA was not investigated<br />

directly.<br />

A second possible mechanism of the preventive role of α-LA is its effects<br />

on lipids in skeletal muscles and pancreatic β cells. Skeletal muscle is the major<br />

tissue responsible for 70 to 80% of whole body glucose uptake. Indeed, insulin<br />

resistance in skeletal muscle is a common characteristic of type 2 diabetes, and<br />

many previous studies in animal models and human subjects have shown that<br />

lipid accumulation in skeletal muscle is associated with insulin resistance in<br />

obesity. In addition, lipid accumulation in pancreatic β cells also affects β cell<br />

damage. LCAC accumulation in pancreatic β cells may induce apoptosis of the<br />

cells. Increased plasma free fatty acid by lipid–heparin infusion induced β cell<br />

hypertrophy. Since α-LA reduces visceral fat mass and decreases plasma free<br />

fatty acid and triglyceride concentrations and tissue triglyceride contents in<br />

OLETF rats, insulin resistance in skeletal muscle and β cell apoptosis can be<br />

prevented.<br />

While the possible mechanisms of the effects of α-LA on the development<br />

of diabetes in obese rats could be inferred from this report, we cannot exclude<br />

the possibility that α-LA prevented the development of diabetes due to its weight<br />

reducing effects. In addition, the exact molecular mechanism by which α-LA<br />

reduces lipid contents in skeletal muscle was not investigated in this report. Thus,<br />

other studies 15,52 were performed to minimize the effect of α-LA on body weight<br />

change and investigate the mechanism of α-LA in reducing lipid content in<br />

skeletal muscle. The authors fed rats for a short period (200 mg/kg body weight)<br />

via intravenous infusion for 2 hr or 0.5% (wt/wt) racemic α-LA (mixed in food<br />

for 3 days) to minimize the effect of α-LA on body weight change. In addition,<br />

since AMPK was known to be responsible for the effect of α-LA in the hypothalamus<br />

and because it increases both glucose uptake and fatty acid oxidation,<br />

the authors focused on AMPK as a molecular target of α-LA. AMPK is known<br />

as a cellular “energy sensor” because its activity is sensitively changed by its<br />

cellular energy state. AMPK controls a number of metabolic processes to<br />

help restore energy depletion in peripheral tissues. 53,54 For example, AMPK activation<br />

in skeletal muscle enhanced glucose uptake and mitochondrial fatty acid


Diabetes Mellitus and Endothelial Dysfunction 265<br />

oxidation. 55 AMPK is also expressed in vascular endothelium, 56 and AMPK<br />

dysregulation has been suggested to contribute to endothelial dysfunction. 57<br />

Administration of α-LA to OLETF rats improved insulin-stimulated whole<br />

body glucose uptake and whole body glycogen synthesis. Insulin-stimulated<br />

glucose uptake and glycogen synthesis in skeletal muscle were also improved in<br />

OLETF rats treated with α-LA. In skeletal muscles of OLETF rats, AMPK (α2-<br />

AMPK, not α1-AMPK) activity was decreased compared to control rats. The<br />

reason for the decreased AMPK activity is not yet known. However, because<br />

AMPK is enzyme activated when the cellular energy state is low, abundant fuels<br />

such as elevated plasma free fatty acid and/or glucose in OLETF rats may reduce<br />

AMPK activity.<br />

Administration of α-LA increased α2-AMPK and fatty acid oxidation and<br />

decreased triglyceride contents in skeletal muscles of OLETF rats. Overexpression<br />

of the dominant negative α2-AMPK gene in skeletal muscles of OLETF rats<br />

reversed the effects of α-LA on triglyceride contents, fatty acid oxidation, and<br />

insulin-stimulated glucose uptake, suggesting that α-LA exerted its effects via<br />

AMPK activation. Because AMPK is known to increase fatty acid oxidation and<br />

because lipid accumulation in muscle is associated with insulin resistance, we<br />

could deduce from this report that α-LA improved insulin sensitivity by activating<br />

AMPK and increasing fatty acid oxidation, and subsequently by reducing lipid<br />

accumulation in skeletal muscle. However, further study is necessary to explain<br />

the precise molecular mechanism by which α-LA prevents the development of<br />

diabetes and improves insulin sensitivity.<br />

α-LIPOIC ACID PREVENTS ENDOTHELIAL DYSFUNCTION IN DIABETES-<br />

PRONE OBESE RATS<br />

Oxidative stress and impaired bioactivity of vascular nitric oxide (NO) play<br />

important roles in the pathogenesis of microvascular and macrovascular complications<br />

in diabetes mellitus. Antioxidative properties and the capacity of increasing<br />

NO synthesis and activity may contribute to the protective role of α-LA in<br />

the development of vascular complications in diabetes. It was reported that O 2 –<br />

production was increased in aortic tissues of insulin-resistant rats, 43,58 and α-LA<br />

supplementation in chronically glucose-fed rats prevents the increase in aortic<br />

basal O 2 – production. In addition, α-LA treatment prevented increases of thiobarbituric<br />

acid-reactive substances (indirect evidence of intensified free radical production<br />

and uniformly increased substances in streptozotocin (STZ)-induced diabetic<br />

rats). 27,28<br />

α-LA improved NO-mediated vasodilation in diabetic patients but not in<br />

healthy control subjects. This effect seems to be associated with antioxidative<br />

properties of α-LA because the effects of α-LA were positively related to plasma<br />

levels of malondialdehyde. 59 In an in vitro study, however, α-LA increased NO<br />

synthesis and bioactivity in human aortic endothelial cells by mechanisms that<br />

appear to be independent of antioxidative properties, because cellular GSH levels<br />

and GSH:GSSG ratios were not significantly changed by α-LA treatment. 60


266 Oxidative Stress and Inflammatory Mechanisms<br />

α-LA improved impaired endothelium-dependent vasorelaxation in diabetic<br />

rats. 61 In STZ-induced diabetic rats, hyperglycemia induces decreased eNOS<br />

expression and increased iNOS expression in tissues such as heart, aorta, sciatic<br />

nerve, and kidney. α-LA treatment increased eNOS expression and decreased<br />

iNOS expression. 62 In addition, the decline in eNOS phosphorylation — a possible<br />

mechanism involved in vascular dysfunction in aging — can be partially restored<br />

by treating old rats with α-LA. 63<br />

Moreover, α-LA treatment inhibited adhesion molecule expression in HAEC<br />

and TNF-α-induced monocyte adhesion. 64 α-LA also inhibited NF-κB-mediated<br />

transcription and expression of endothelial genes such as tissue factor and endothelin-1<br />

65 and reduced advanced glycation end product (AGE)-induced endothelial<br />

expression of VCAM-1 and monocyte binding to the endothelium. 66<br />

In addition to these mechanisms, α-LA may exhibit diverse protective effects<br />

in vascular disease. It may lower blood pressure in rats. α-LA supplementation<br />

decreased systolic blood pressure in spontaneously hypertensive rats, 67 in high<br />

salt-treated rats, 68 and in fructose-induced hypertensive WKY rats. 69 Similarly,<br />

hypertension induced by the addition of a 10% D-glucose drink to their diet was<br />

attenuated by α-LA in Sprague-Dawley rats. 70<br />

α-LA has exhibited lipid-lowering capacity. In studies of rabbits 71,72 and<br />

Japanese quail, 73 α-LA decreased levels of cholesterol and lipoprotein in serum.<br />

It also reduced serum triglyceride levels in STZ-induced diabetic rats. 24<br />

α-LA has been shown to decrease LDL oxidation. It was reported that DHLA<br />

inhibited Cu 2+ -dependent LDL peroxidation by chelating copper in in vitro experiments.<br />

74 In addition, in human studies, α-LA (600 mg/day) significantly<br />

increased the lag time of LDL lipid peroxide formation for both copper-catalyzed<br />

and 2,2′-azobis (2-amidinopropane) hydrochloride (AAPH)-induced LDL oxidation.<br />

75<br />

Central obesity is an important risk factor for cardiovascular disease. Endothelial<br />

dysfunction, generally considered a prerequisite for atherosclerosis, is a<br />

frequent finding in obesity. A recent report indicates that administration of α-LA<br />

to patients with metabolic syndrome improved endothelial function 76 but the<br />

mechanism is not yet known. It was suggested that lipid accumulation in vascular<br />

tissue and a consequent increase in oxidative stress lead to endothelial dysfunction<br />

in obesity. 39<br />

As described in the previous section, α-LA enhances fatty acid oxidation and<br />

reduces lipid accumulation by activating AMPK in skeletal muscles of obese rats.<br />

Because α-LA produces these effects and because AMPK is expressed in vascular<br />

endothelial cells, 77 it is conceivable that α-LA may improve endothelial function<br />

in obesity by activating AMPK. One recent study 15 demonstrated that endothelium-dependent<br />

vasorelaxation was impaired and AMPK activity in endothelium<br />

of obese animals was decreased compared to control (lean) rats. In obese rats,<br />

α-LA administration improved the impaired endothelium-dependent vasorelaxation<br />

and increased AMPK activity and phosphorylation independent of the effect<br />

of α-LA on body weight. In addition, in an in vitro study using human aortic<br />

endothelial cells (HAECs) treated with a fatty acid (linoleic acid), α-LA prevented


Diabetes Mellitus and Endothelial Dysfunction 267<br />

linoleic acid-induced decreases in AMPK phosphorylation. This effect was associated<br />

with normalization of endothelial apoptosis and ROS generation in the<br />

presence of linoleic acid. These results suggest that the mechanism of endothelial<br />

dysfunction in obesity is reduced AMPK activity in endothelial cells, and that α-<br />

LA may improve endothelial dysfunction in obese rats by activating AMPK in<br />

endothelial cells.<br />

SUMMARY<br />

A growing body of evidence suggests that α-LA exerts beneficial effects on both<br />

prevention and treatment of oxidative stress in a number of models and clinical<br />

conditions including diabetes and vascular dysfunction. However, before α-LA<br />

can be used in clinical practice to prevent diabetes or vascular dysfunction, more<br />

experimental research to elicit the mechanism and clinical trials to determine the<br />

proper dose, formula, and route of administration are needed.<br />

ACKNOWLEDGMENT<br />

This work was supported by National Research Laboratory grant<br />

M1040000000804J000000810 from the Ministry of Science and Technology of<br />

the Republic of Korea.<br />

REFERENCES<br />

1. Jones, W., Li, X., Qu, Z.C., Perriott, L., Whitesell, R.R., and May, J.M. (2002)<br />

Uptake, recycling, and antioxidant actions of alpha-lipoic acid in endothelial cells.<br />

Free Radic Biol Med 33, 83.<br />

2. Packer, L., Roy, S., and Sen, C.K. (1997) Alpha-lipoic acid: a metabolic antioxidant<br />

and potential redox modulator of transcription. Adv Pharmacol 38, 79.<br />

3. Reed, L.J. (1998) From lipoic acid to multi-enzyme complexes. Protein Sci 7, 220.<br />

4. Biewenga, G.P., Haenen, G.R., and Bast, A. (1997) The pharmacology of the<br />

antioxidant lipoic acid. Gen Pharmacol 29, 315.<br />

5. Packer, L., Witt, E.H., and Tritschler, H.J. (1995) Alpha-lipoic acid as a biological<br />

antioxidant. Free Radic Biol Med 19, 227.<br />

6. Kramer, K.P., Hoppe, P., and Packer, L., Eds. (2001) Nutraceuticals in Health and<br />

Disease Prevention, Marcel Dekker, New York, p. 129.<br />

7. Obrosova, I., Cao, X., Greene, D.A., and Stevens, M.J. (1998) Diabetes-induced<br />

changes in lens antioxidant status, glucose utilization and energy metabolism:<br />

effect of DL-alpha-lipoic acid. Diabetologia 41, 1442.<br />

8. Scott, B.C. et al. (1994) Lipoic and dihydrolipoic acids as antioxidants: a critical<br />

evaluation. Free Radic Res 20, 119.<br />

9. Packer, L. (1998) Alpha-lipoic acid: a metabolic antioxidant which regulates NFkappa<br />

B signal transduction and protects against oxidative injury. Drug Metab<br />

Rev 30, 245.


268 Oxidative Stress and Inflammatory Mechanisms<br />

10. Evans, J.L. and Goldfine, I.D. (2000) Alpha-lipoic acid: a multifunctional antioxidant<br />

that improves insulin sensitivity in patients with type 2 diabetes. Diabetes<br />

Technol Ther 2, 401.<br />

11. Suzuki, Y.J., Tsuchiya, M., and Packer, L. (1992) Lipoate prevents glucose-induced<br />

protein modifications. Free Radic Res Commun 17, 211.<br />

12. Roy, S., Sen, C.K., Tritschler, H.J., and Packer, L. (1997) Modulation of cellular<br />

reducing equivalent homeostasis by alpha-lipoic acid: mechanisms and implications<br />

for diabetes and ischemic injury. Biochem Pharmacol 53, 393.<br />

13. Romero, F.J., Ordonez, I., Arduini, A., and Cadenas, E. (1992) The reactivity of<br />

thiols and disulfides with different redox states of myoglobin: redox and addition<br />

reactions and formation of thiyl radical intermediates. J Biol Chem 267, 1680.<br />

14. Song, K.H. et al. (2005) Alpha-lipoic acid prevents diabetes mellitus in diabetesprone<br />

obese rats. Biochem Biophys Res Commun 326, 197.<br />

15. Lee, W.J. et al. (2005) Alpha-lipoic acid prevents endothelial dysfunction in obese<br />

rats via activation of AMP-activated protein kinase. Arterioscler Thromb Vasc Biol<br />

25, 2488.<br />

16. Packer, L., Tritschler, H.J., and Wessel, K. (1997) Neuroprotection by the metabolic<br />

antioxidant alpha-lipoic acid. Free Radic Biol Med 22, 359.<br />

17. Ceriello, A. (2000) Oxidative stress and glycemic regulation. Metabolism 49, 27.<br />

18. Baynes, J.W. and Thorpe, S.R. (1999) Role of oxidative stress in diabetic complications:<br />

a new perspective on an old paradigm. Diabetes 48, 1.<br />

19. Baynes, J.W. (1991) Role of oxidative stress in development of complications in<br />

diabetes. Diabetes 40, 405.<br />

20. Chang, K.C. et al. (1993) Possible superoxide radical-induced alteration of vascular<br />

reactivity in aortas from streptozotocin-treated rats. J Pharmacol Exp Ther<br />

266, 992.<br />

21. Young, I.S., Tate, S., Lightbody, J.H., McMaster, D., and Trimble, E.R. (1995)<br />

The effects of desferrioxamine and ascorbate on oxidative stress in the streptozotocin<br />

diabetic rat. Free Radic Biol Med 18, 833.<br />

22. Saxena, A.K., Srivastava, P., Kale, R.K., and Baquer, N.Z. (1993) Impaired antioxidant<br />

status in diabetic rat liver: effect of vanadate. Biochem Pharmacol 45, 539.<br />

23. McLennan, S.V. et al. (1991) Changes in hepatic glutathione metabolism in diabetes.<br />

Diabetes 40, 344.<br />

24. Ford, I., Cotter, M.A., Cameron, N.E., and Greaves, M. (2001) The effects of<br />

treatment with alpha-lipoic acid or evening primrose oil on vascular hemostatic<br />

and lipid risk factors, blood flow, and peripheral nerve conduction in the streptozotocin-diabetic<br />

rat. Metabolism 50, 868.<br />

25. Kishi, Y. et al. (1999) Alpha-lipoic acid: effect on glucose uptake, sorbitol pathway,<br />

and energy metabolism in experimental diabetic neuropathy. Diabetes 48, 2045.<br />

26. Ziegler, D. and Gries, F.A. (1997) Alpha-lipoic acid in the treatment of diabetic<br />

peripheral and cardiac autonomic neuropathy. Diabetes 46, Suppl 2, S62.<br />

27. Obrosova, I.G., Fathallah, L., and Greene, D.A. (2000) Early changes in lipid<br />

peroxidation and antioxidative defense in diabetic rat retina: effect of DL-alphalipoic<br />

acid. Eur J Pharmacol 398, 139.<br />

28. Kocak, G. et al. (2000) Alpha-lipoic acid treatment ameliorates metabolic parameters,<br />

blood pressure, vascular reactivity and morphology of vessels already damaged<br />

by streptozotocin-diabetes. Diabetes Nutr Metab 13, 308.


Diabetes Mellitus and Endothelial Dysfunction 269<br />

29. Morcos, M. et al. (2001) Effect of alpha-lipoic acid on the progression of endothelial<br />

cell damage and albuminuria in patients with diabetes mellitus: an exploratory<br />

study. Diabetes Res Clin Pract 52, 175.<br />

30. Jacob, S. et al. (1996) The antioxidant alpha-lipoic acid enhances insulin-stimulated<br />

glucose metabolism in insulin-resistant rat skeletal muscle. Diabetes 45,<br />

1024.<br />

31. Henriksen, E.J. et al. (1997) Stimulation by alpha-lipoic acid of glucose transport<br />

activity in skeletal muscle of lean and obese Zucker rats. Life Sci 61, 805.<br />

32. Jacob, S. et al. (1995) Enhancement of glucose disposal in patients with type 2<br />

diabetes by alpha-lipoic acid. Arzneimittelforschung 45, 872.<br />

33. Jacob, S., Henriksen, E.J., Tritschler, H.J., Augustin, H.J., and Dietze, G.J. (1996)<br />

Improvement of insulin-stimulated glucose-disposal in type 2 diabetes after<br />

repeated parenteral administration of thioctic acid. Exp Clin Endocrinol Diabetes<br />

104, 284.<br />

34. Estrada, D.E. et al. (1996) Stimulation of glucose uptake by the natural coenzyme<br />

alpha-lipoic acid/thioctic acid: participation of elements of the insulin signaling<br />

pathway. Diabetes 45, 1798.<br />

35. Haugaard, N. and Haugaard, E.S. (1970) Stimulation of glucose utilization by<br />

thioctic acid in rat diaphragm incubated in vitro. Biochim Biophys Acta 222, 583.<br />

36. Singh, H.P. and Bowman, R.H. (1970) Effect of DL-alpha-lipoic acid on the citrate<br />

concentration and phosphofructokinase activity of perfused hearts from normal<br />

and diabetic rats. Biochem Biophys Res Commun 41, 555.<br />

37. Lieberman, L.S. (2003) Dietary, evolutionary, and modernizing influences on the<br />

prevalence of type 2 diabetes. Annu Rev Nutr 23, 345.<br />

38. Unger, R.H. and Zhou, Y.T. (2001) Lipotoxicity of beta-cells in obesity and in<br />

other causes of fatty acid spillover. Diabetes 50, Suppl 1, S118.<br />

39. Bakker, S.J. et al. (2000) Cytosolic triglycerides and oxidative stress in central<br />

obesity: the missing link between excessive atherosclerosis, endothelial dysfunction,<br />

and beta-cell failure? Atherosclerosis 148, 17.<br />

40. Evans, J.L., Goldfine, I.D., Maddux, B.A., and Grodsky, G.M. (2002) Oxidative<br />

stress and stress-activated signaling pathways: a unifying hypothesis of type 2<br />

diabetes. Endocr Rev 23, 599.<br />

41. Rudich, A., Tirosh, A., Potashnik, R., Khamaisi, M., and Bashan, N. (1999) Lipoic<br />

acid protects against oxidative stress induced impairment in insulin stimulation<br />

of protein kinase B and glucose transport in 3T3-L1 adipocytes. Diabetologia 42,<br />

949.<br />

42. Maddux, B.A., See, W., Lawrence, J.C., Jr., Goldfine, A.L., Goldfine, I.D., and<br />

Evans, J.L. (2001) Protection against oxidative stress-induced insulin resistance<br />

in rat L6 muscle cells by mircomolar concentrations of alpha-lipoic acid. Diabetes<br />

50, 404.<br />

43. El Midaoui, A. and de Champlain, J. (2002) Prevention of hypertension, insulin<br />

resistance, and oxidative stress by alpha-lipoic acid. Hypertension 39, 303.<br />

44. Jacob, S. et al. (1999) Oral administration of RAC-alpha-lipoic acid modulates<br />

insulin sensitivity in patients with type-2 diabetes mellitus: a placebo-controlled<br />

pilot trial. Free Radic Biol Med 27, 309.<br />

45. Eason, R.C., Archer, H.E., Akhtar, S., and Bailey, C.J. (2002) Lipoic acid increases<br />

glucose uptake by skeletal muscles of obese-diabetic ob/ob mice. Diabetes Obes<br />

Metab 4, 29.


270 Oxidative Stress and Inflammatory Mechanisms<br />

46. Khamaisi, M. et al. (1999) Lipoic acid acutely induces hypoglycemia in fasting<br />

nondiabetic and diabetic rats. Metabolism 48, 504.<br />

47. Khamaisi, M. et al. (1997) Lipoic acid reduces glycemia and increases muscle<br />

GLUT4 content in streptozotocin-diabetic rats. Metabolism 46, 763.<br />

48. Kim, M.S. et al. (2004) Anti-obesity effects of alpha-lipoic acid mediated by<br />

suppression of hypothalamic AMP-activated protein kinase. Nat Med 10, 727.<br />

49. Kawano, K., Hirashima, T., Mori, S., Saitoh, Y., Kurosumi, M., and Natori, T.<br />

(1992) Spontaneous long-term hyperglycemic rat with diabetic complications:<br />

Otsuka Long-Evans Tokushima fatty (OLETF) strain. Diabetes 41, 1422.<br />

50. Hansen, L.L., Ikeda, Y., Olsen, G.S., Busch, A.K., and Mosthaf, L. (1999) Insulin<br />

signaling is inhibited by micromolar concentrations of H 2O 2: evidence for a role<br />

of H 2O 2 in tumor necrosis factor alpha-mediated insulin resistance. J Biol Chem<br />

274, 25078.<br />

51. Tirosh, A., Potashnik, R., Bashan, N., and Rudich, A. (1999) Oxidative stress<br />

disrupts insulin-induced cellular redistribution of insulin receptor substrate-1 and<br />

phosphatidylinositol 3-kinase in 3T3-L1 adipocytes: a putative cellular mechanism<br />

for impaired protein kinase B activation and GLUT4 translocation. J Biol Chem<br />

274, 10595.<br />

52. Lee, W.J. et al. (2005) Alpha-lipoic acid increases insulin sensitivity by activating<br />

AMPK in skeletal muscle. Biochem Biophys Res Commun 332, 885.<br />

53. Kemp, B.E. et al. (2003) AMP-activated protein kinase, super metabolic regulator.<br />

Biochem Soc Trans 31, 162.<br />

54. Hardie, D.G., Salt, I.P., Hawley, S.A., and Davies, S.P. (1999) AMP-activated<br />

protein kinase: an ultrasensitive system for monitoring cellular energy charge.<br />

Biochem J 338 (Pt 3), 717.<br />

55. Kahn, B.B., Alquier, T., Carling, D., and Hardie, D.G. (2005) AMP-activated<br />

protein kinase: ancient energy gauge provides clues to modern understanding of<br />

metabolism. Cell Metab 1, 15.<br />

56. Dagher, Z., Ruderman, N., Tornheim, K., and Ido, Y. (2001) Acute regulation of<br />

fatty acid oxidation and amp-activated protein kinase in human umbilical vein<br />

endothelial cells. Circ Res 88, 1276.<br />

57. Ruderman, N. and Prentki, M. (2004) AMP kinase and malonyl-CoA: targets for<br />

therapy of the metabolic syndrome. Nat Rev Drug Discov 3, 340.<br />

58. Kashiwagi, A. et al. (1999) Free radical production in endothelial cells as a<br />

pathogenetic factor for vascular dysfunction in the insulin resistance state. Diabetes<br />

Res Clin Pract 45, 199.<br />

59. Heitzer, T. et al. (2001) Beneficial effects of alpha-lipoic acid and ascorbic acid<br />

on endothelium-dependent, nitric oxide-mediated vasodilation in diabetic patients:<br />

relation to parameters of oxidative stress. Free Radic Biol Med 31, 53.<br />

60. Visioli, F., Smith, A., Zhang, W., Keaney, J.F., Jr., Hagen, T., and Frei, B. (2002)<br />

Lipoic acid and vitamin C potentiate nitric oxide synthesis in human aortic endothelial<br />

cells independently of cellular glutathione status. Redox Rep 7, 223.<br />

61. Cameron, N.E., Jack, A.M., and Cotter, M.A. (2001) Effect of alpha-lipoic acid<br />

on vascular responses and nociception in diabetic rats. Free Radic Biol Med 31,<br />

125.<br />

62. Bojunga, J., Dresar-Mayert, B., Usadel, K.H., Kusterer, K., and Zeuzem, S. (2004)<br />

Antioxidative treatment reverses imbalances of nitric oxide synthase isoform<br />

expression and attenuates tissue-cGMP activation in diabetic rats. Biochem Biophys<br />

Res Commun 316, 771.


Diabetes Mellitus and Endothelial Dysfunction 271<br />

63. Smith, A.R. and Hagen, T.M. (2003) Vascular endothelial dysfunction in aging:<br />

loss of Akt-dependent endothelial nitric oxide synthase phosphorylation and partial<br />

restoration by (R)-alpha-lipoic acid. Biochem Soc Trans 31, 1447.<br />

64. Zhang, W.J. and Frei, B. (2001) Alpha-lipoic acid inhibits TNF-alpha-induced<br />

NF-kappaB activation and adhesion molecule expression in human aortic endothelial<br />

cells. FASEB J 15, 2423.<br />

65. Bierhaus, A. et al. (1997) Advanced glycation end product-induced activation of<br />

NF-kappa-B is suppressed by alpha-lipoic acid in cultured endothelial cells. Diabetes<br />

46, 1481.<br />

66. Kunt, T. et al. Alpha-lipoic acid reduces expression of vascular cell adhesion<br />

molecule-1 and endothelial adhesion of human monocytes after stimulation with<br />

advanced glycation end products. Clin Sci (Lond) 96, 75.<br />

67. Vasdev, S., Ford, C.A., Parai, S., Longerich, L., and Gadag, V. (2000) Dietary<br />

alpha-lipoic acid supplementation lowers blood pressure in spontaneously hypertensive<br />

rats. J Hypertens 18, 567.<br />

68. Vasdev, S., Gill, V., Longerich, L., Parai, S., and Gadag, V. (2003) Salt-induced<br />

hypertension in WKY rats: prevention by alpha-lipoic acid supplementation. Mol<br />

Cell Biochem 254, 319.<br />

69. Vasdev, S., Ford, C.A., Parai, S., Longerich, L., and Gadag, V. (2000) Dietary<br />

lipoic acid supplementation prevents fructose-induced hypertension in rats. Nutr<br />

Metab Cardiovasc Dis 10, 339.<br />

70. Midaoui, A.E., Elimadi, A., Wu, L., Haddad, P.S., and de Champlain, J. (2003)<br />

Lipoic acid prevents hypertension, hyperglycemia, and the increase in heart mitochondrial<br />

superoxide production. Am J Hypertens 16, 173.<br />

71. Angelucci, L. and Mascitelli-Coriandoli, E. (1958) Anticholesterol activity of<br />

alpha-lipoic acid. Nature 181, 911.<br />

72. Ivanov, V.N. (1974) Effect of lipoic acid on tissue respiration in rabbits with<br />

experimental atherosclerosis. Cor Vasa 16, 141.<br />

73. Shih, J.C. (1983) Atherosclerosis in Japanese quail and the effect of lipoic acid.<br />

Fed Proc 42, 2494.<br />

74. Lodge, J.K., Traber, M.G., and Packer, L. (1998) Thiol chelation of Cu 2+ by<br />

dihydrolipoic acid prevents human low density lipoprotein peroxidation. Free<br />

Radic Biol Med 25, 287.<br />

75. Marangon, K., Devaraj, S., Tirosh, O., Packer, L., and Jialal, I. (1999) Comparison<br />

of the effect of alpha-lipoic acid and alpha-tocopherol supplementation on measures<br />

of oxidative stress. Free Radic Biol Med 27, 1114.<br />

76. Sola, S. et al. (2005) Irbesartan and lipoic acid improve endothelial function and<br />

reduce markers of inflammation in the metabolic syndrome: results of the Irbesartan<br />

and Lipoic Acid in Endothelial Dysfunction (ISL<strong>AND</strong>) study. Circulation<br />

111, 343.<br />

77. Dagher, Z., Ruderman, N., Tornheim, K., and Ido, Y. (1999) The effect of AMPactivated<br />

protein kinase and its activator AICAR on the metabolism of human<br />

umbilical vein endothelial cells. Biochem Biophys Res Commun 265, 112.


18<br />

CONTENTS<br />

Lipoic Acid Blocks<br />

Obesity through<br />

Reduced Food Intake,<br />

Enhanced Energy<br />

Expenditure, and<br />

Inhibited Adipocyte<br />

Differentiation<br />

Jong-Min Park and An-Sik Chung<br />

Introduction .......................................................................................................274<br />

Obesity Related to Fatty Cell Biology, Energy Expenditure, and<br />

Feeding Regulation..................................................................................274<br />

Antiobesity Function through Suppression of Hypothalamic AMPK .............277<br />

Regulation of Energy Expenditure and Food Intake ..............................277<br />

Role of Hypothalamic AMPK.................................................................278<br />

Reduction of Food Intake and Increase of Energy Expenditure via<br />

Suppression of Hypothalamic AMPK.........................................278<br />

Regulation of Adipocyte Differentiation ..........................................................280<br />

Inhibition of Adipocyte Differentiation via Mitogen-Activated Protein<br />

Kinase Pathways......................................................................................281<br />

MAPK Signaling Pathways Mediate Actions of LA<br />

on Adipogenesis ..........................................................................281<br />

Modulation of Auxiliary Transcription Factors in Adipogenesis...........281<br />

MAPK Signaling Pathways Mediate LA Actions on Cell<br />

Cycle and Clonal Expansion.......................................................282<br />

Summary ...........................................................................................................283<br />

Acknowledgments .............................................................................................283<br />

References .........................................................................................................284<br />

273


274 Oxidative Stress and Inflammatory Mechanisms<br />

INTRODUCTION<br />

Obesity is rapidly increasing throughout the world, substantially shortening life<br />

expectancy, and closely correlated with the prevalence of diabetes and cardiovascular<br />

diseases. Plasma levels of leptin, tumor necrosis factor (TNF)-α and nonesterified<br />

fatty acids are elevated in obesity and substantially contribute to the<br />

development of insulin resistance. 1<br />

Although obesity research including studies of leptin 2 and the leptin receptor<br />

gene 3 has been extensively pursued for more than two decades, the molecular<br />

mechanisms of obesity are not yet completely understood. Finding target molecules<br />

of weight regulatory mechanisms will contribute to the development of safe<br />

and effective pharmaceuticals for blocking obesity and preventing diabetes and<br />

cardiovascular diseases. α-lipoic acid (LA) and its reduced dihydrolipoic acid<br />

(DHLA) are considered antioxidants (Figure 18.1). LA has components of αketo<br />

dehydrogenases including pyruvate dehydrogenases that catalyze various<br />

redox-based reactions. 4 Recent studies demonstrated that LA facilitates glucose<br />

transport and utilization in fully differentiated adipocytes as well as animal models<br />

of diabetes. 5–7 LA also dramatically reduced rodent weights by suppressing food<br />

intake and increasing energy expenditures. 8<br />

Obesity is caused by both hypertrophy of adipose tissue and by adipose tissue<br />

hyperplasia, which triggers the transformation of pre-adipocytes into adipocytes. 9<br />

LA decreases hypothalamic AMP-activated protein kinase (AMPK) activity and<br />

induces preformed weight losses in rodents by reducing food intake and enhancing<br />

energy expenditures. 8 This chapter will focus on antiobesity mechanisms of<br />

LA: blocking adipocyte differentiation and reducing hypothalamic AMPK.<br />

OBESITY RELATED TO FATTY CELL BIOLOGY, ENERGY<br />

EXPENDITURE, <strong>AND</strong> FEEDING REGULATION<br />

Adipose tissue is the major storage organ for surplus energy. It is now clear that<br />

adipose tissue is a complex and highly active metabolic and endocrine organ.<br />

S<br />

S<br />

*<br />

COOH<br />

α-Lipoic acid<br />

(1,2-dithiolane-3-pentanoic acid)<br />

Reduction<br />

(2H + +2e - )<br />

SH SH<br />

FIGURE 18.1 Structures of α-lipoic acid and dihyrolipoic acid.<br />

E<br />

E: Lipoamide dehydrogenases<br />

Glutathione reductases<br />

Thioredoxin reductases<br />

*<br />

COOH<br />

Dihydrolipoic acid (DHLA)


Lipoic Acid Blocks Obesity 275<br />

Leptin is a representative adipocyte-derived hormone that signals information<br />

about the body’s energy status from the adipose tissue to the brain.<br />

Although it is well known that leptin-deficient ob/ob and leptin receptor<br />

deficient db/db mice develop severe obesity, leptin deficiency is very rare in<br />

humans. Most obese individuals have increased plasma leptin concentrations,<br />

suggesting that they are leptin-resistant. In addition to secreting leptin, adipose<br />

tissue secretes a variety of bioactive peptides, collectively called adipocytokines,<br />

including TNF-α, adiponectin, plasminogen activator inhibitor (PAI)-1, interleukin<br />

(IL)-6, and angiotensinogen. The expression profiles of adipocytokines in<br />

subcutaneous adipose tissues and those in visceral adipose tissues are different.<br />

Adiponectin, PAI-1, IL-6, and angiotensinogen are mainly shown in the latter.<br />

The change in these important adipocytokines by visceral obesity is regarded<br />

as the cause of the detrimental metabolic effects. Recently, the physiologic role<br />

of adiponectin has received considerable attention. Adiponectin has been shown<br />

to reduce insulin resistance and atherosclerotic processes and to increase fatty<br />

acid oxidation rates. 10 The actions of adiponectin are considered to arise from<br />

the activation of AMPK through the recently cloned adiponectin receptor. 11 Unlike<br />

other adipocytokines, plasma adiponectin levels are reduced in accordance with<br />

body (visceral) fat mass. 10 TNF-α, which increases in obesity and inhibits insulin<br />

signals, is considered a key factor in the regulation of adiponectin production.<br />

Body weight results from a balance between food intake and energy expenditure.<br />

The sympathetic nervous system is activated in response to excess energy<br />

or a cold environment. Mice with deletions of the genes encoding the three<br />

adrenergic receptor subtypes 12 developed severe obesity as a result of their<br />

inability to increase energy expenditure in response to a high caloric diet. In<br />

rodents, uncoupling protein (UCP)-1 in brown adipose tissue (BAT) is the main<br />

regulator of basal energy expenditure and the expression of this protein is<br />

increased by adrenergic stimulation. In humans, however, the main regulator of<br />

energy expenditure is not yet known. The UCP-1 homologues known as UCP-2<br />

and UCP-3 that are expressed in human tissues were formerly thought to be the<br />

main regulators of energy expenditure. Indeed, hyperexpression of UCP-2 and<br />

UCP-3 was shown to increase energy expenditure and reduce body fat. These<br />

proteins, however, may not be the major regulators of whole body energy expenditure,<br />

inasmuch as mice deficient in either protein did not develop obesity. 13<br />

The neural system that regulates body weight and appetite is centered in the<br />

hypothalamus, which coordinates both afferent sensing and efferent action signals.<br />

Long-term afferent signals such as leptin and insulin sense the long-term<br />

status of body energy stores, whereas short-term (meal-related) afferent signals<br />

derived from the gut are involved in regulating the onset or termination of<br />

individual meals. Neuronal cells, which sense nutrient availability, trigger feeding<br />

behavior. 13<br />

Intracerebroventricular (ICV) administration of glucose or long chain fatty<br />

acid has been found to inhibit food intake, 14 whereas central administration of 2deoxyglucose<br />

(2-DG), a non-metabolizable glucose analogue, or mercaptoacetate,<br />

an inhibitor of fatty acid oxidation, elicits feeding behavior. 15


276 Oxidative Stress and Inflammatory Mechanisms<br />

Leptin, insulin, glucose, α-lipoic acid<br />

AMPK↓<br />

Acetyl-CoA<br />

ACC↑<br />

CPT-1↓<br />

β-oxidation↓<br />

Hypothalamus<br />

Malonyl-CoA↑<br />

LCAC↑<br />

Food intake↓<br />

FIGURE 18.2 Possible mechanism by which decreased hypothalamic AMP-activated protein<br />

kinase (AMPK) activity reduces food intake. The reduction in hypothalamic AMPK<br />

activity in response to feeding-inhibiting factors such as leptin, insulin, glucose, and α-<br />

LA increases acetyl-CoA carboxylase (ACC) activity. Increased ACC activity leads to<br />

increase in malonyl-CoA levels, which in turn inhibits carnitine palmitoyltransferase-1<br />

(CPT-1) and mitochondrial β oxidation of long chain fatty acyl-CoA (LCAC). Recent<br />

studies suggest that increases in malonyl-CoA and/or LCAC levels in the hypothalamus<br />

decrease food intake. 16,18,28<br />

Several signaling pathways are thought to be involved in mediating nutrientinduced<br />

feeding. For example, central administration of the fatty acid synthase<br />

(FAS) inhibitors, cerulenin and C75, reduces food intake, but it can be prevented<br />

by the coadministration of the acetyl-CoA carboxylase (ACC) inhibitor, TOFA. 16<br />

This result suggests that malonyl-CoA, an intermediate metabolite between ACC<br />

and FAS, may be an anorexigenic signal (Figure 18.2).<br />

Patients with metabolic syndrome (MS) are characterized by insulin resistance,<br />

obesity, hyperlipidemia, premature atherosclerosis, and type-2 diabetes. It<br />

has been proposed that the common feature linking this syndrome is dysregulation<br />

of the AMPK/malonyl-CoA fuel-sensing and signaling network. 17 Both fuel surfeit<br />

and reduced AMPK activity result in increased cellular malonyl-CoA levels,<br />

which reduces fat oxidation and favors abnormal tissue accumulation of lipids.<br />

Some evidential factors indicate that activated AMPK and/or reduced malonyl-<br />

CoA levels may reverse these abnormalities or prevent them from occurring. In<br />

contrast, inhibition of hypothalamic carnitine palmitoyl transferase-1 (CPT-1)<br />

reduces food intake. 18<br />

These findings led to the suggestion that increased cytosolic concentrations<br />

of long chain fatty acyl-CoA and malonyl-CoA levels may serve as anorexigenic<br />

signals. Like pancreatic β cells, some neurons in the ventromedial and arcuate<br />

nuclei of the hypothalamus have glucose-sensing machinery that includes<br />

GLUT2, glucokinase, and the ATP-sensitive potassium (KATP) channel. 19 The


Lipoic Acid Blocks Obesity 277<br />

anorexic hormones leptin and insulin can activate the KATP channel in glucoseresponsive<br />

hypothalamic neurons. 20<br />

ANTIOBESITY FUNCTION THROUGH SUPPRESSION OF<br />

HYPOTHALAMIC AMPK<br />

REGULATION OF ENERGY EXPENDITURE <strong>AND</strong> FOOD INTAKE<br />

Chronic LA treatment significantly reduced body weight gain and visceral fat<br />

mass in obese Otsuta Long-Evans Tokushima fatty (OLETF) rats. 21 Male Sprague-<br />

Dawley rats given a standard chow containing LA for 2 weeks reduced food<br />

intake and body weight in a dose-dependent manner. Acute administration of LA<br />

by intraperitoneal injection (50, 75, or 100 mg/kg of body weight) also suppressed<br />

food intake. In addition, ICV injection of a small amount of LA reduced food<br />

intake, suggesting that the CNS is the primary site of the anorexic effect.<br />

Dietary administration of LA (0.5%, w/w) for 14 weeks also decreased body<br />

weight and visceral fat mass in genetically OLETF rats. This effect was accompanied<br />

by reductions in plasma glucose, insulin, free fatty acids, and leptin. It<br />

has been further demonstrated that LA improved endothelial dysfunction in obese<br />

rats by activating AMPK in endothelial cells by reducing glyceride and lipid<br />

peroxide and increasing NO synthesis. 22<br />

LA also increases whole body energy expenditure. One study compared body<br />

weight changes in rats given dietary LA (0.5%) and in pair-fed rats given the<br />

same amount of food. The LA-treated rats weighed significantly less than the<br />

pair-fed rats, indicating enhanced use of ingested energy. Energy expenditure<br />

measured by indirect calorimetry was higher in rats given LA than in control or<br />

pair-fed rats. UCP-1 in BAT is a chief regulator of energy expenditures in<br />

rodents. 23 UCP-1 is located in the mitochondrial inner membranes and dissipates<br />

proton electrochemical energy as heat. The pair-fed rats show reduced expression<br />

of UCP-1 in BAT and ectopic expression of UCP-1 in white adipose tissue. These<br />

results suggest that weight loss induced by LA is due in part to an enhancement<br />

of energy expenditure.<br />

The effects of LA on food intake and energy metabolism are similar to the<br />

reported effects of leptins 1 and 2. To determine whether the action of LA is<br />

mediated by leptin or leptin receptor signaling, leptin-deficient (Lep–/–) or leptin<br />

receptor-deficient (Lepr–/–) mice were fed a diet containing LA (0.5%). LA<br />

reduced food intake and caused weight losses in both strains of mice, indicating<br />

that leptin and its receptor are not essential for LA-induced anorexia. Leptin<br />

exerts its anorexic effect through the regulation of hypothalamic neuropeptides. 24<br />

Intraperitoneal (i.p.) administration of leptin (1 mg/kg) to Sprague-Dawley rats<br />

reduced the expression of hypothalamic neuropeptide Y (NPY) mRNA and<br />

increased that of pro-opiomelanocortin (Pomc) and corticotropin releasing hormone<br />

(Crh) mRNA. By contrast, the i.p. administration of LA (75 mg/kg) did<br />

not cause any acute changes in the levels of hypothalamic NPY, Pomc, Crh, promelanin<br />

concentrating hormone, or hypocretin mRNA.


278 Oxidative Stress and Inflammatory Mechanisms<br />

ROLE OF HYPOTHALAMIC AMPK<br />

AMPK is an enzyme that acts as an intracellular energy sensor. 25 It is activated<br />

when cell energy status is low. When activated, AMPK inhibits ATP-consuming<br />

pathways (e.g., fatty acid synthesis) and activates ATP-generating pathways (e.g.,<br />

fatty acid oxidation and glycolysis), thus maintaining energy balance within cells.<br />

AMPK activation in skeletal muscle enhances glucose uptake and mitochondrial<br />

fatty acid oxidation. 25<br />

In the liver, AMPK activation suppresses endogenous glucose production. 25<br />

In pancreatic β cells, AMPK seems to antagonize the effect of glucose on insulin<br />

secretion and induce β cell apoptosis. 26 Activation of AMPK has been reported<br />

to play a favorable role in preserving β cell function under lipid overloading<br />

conditions. 27<br />

Recent studies 21,28,29 have demonstrated that AMPK activity in hypothalamic<br />

neurons is altered by various factors and mediates their feeding effects. Hypothalamic<br />

AMPK activity is regulated by nutritional availability in hypothalamic<br />

neurons. Administration of 2-DG increases hypothalamic AMPK activity, while<br />

co-administration of an AMPK inhibitor, compound C, inhibited the 2-DGinduced<br />

glycolysis activity. 21 Conversely, ICV administration of glucose or restoration<br />

of food intake has been found to decrease hypothalamic AMPK activity. 29<br />

AMPK activity is reduced by ICV administration of anorexigenic hormones such<br />

as insulin and leptin, but increased by ICV administration of ghrelin, an orexigenic<br />

hormone. 28,29 In the hypothalamic paraventricular nucleus, AMPK activity is<br />

decreased by the MT-II melanocortin receptor agonist but increased by the melanocortin<br />

receptor antagonist, agouti-related protein (AgRP).<br />

Taken together, these findings indicate that AMPK is part of the common<br />

signaling pathway by which various factors regulate feeding behavior, that is,<br />

hypothalamic AMPK activity is reduced by feeding inhibiting factors and increased<br />

by feeding stimulating factors. Although the mechanism by which AMPK activity<br />

in hypothalamic neurons affects feeding behavior is still not fully understood, the<br />

leptin-induced reduction in hypothalamic AMPK activity is shown to decrease<br />

feeding by down-regulating expression of the NPY and AgRP orexigenic hormones.<br />

29 Alternatively, changes in AMPK activity may affect feeding via changes<br />

in intracellular malonyl-CoA concentrations and CPT-1 activity. 16,18<br />

REDUCTION OF FOOD INTAKE <strong>AND</strong> INCREASE OF ENERGY EXPENDITURE<br />

VIA SUPPRESSION OF HYPOTHALAMIC AMPK<br />

It was recently demonstrated that LA has anti-obesity effects mediated by the<br />

suppression of hypothalamic AMPK activity. 21 LA is an essential cofactor of<br />

mitochondrial pyruvate dehydrogenase and α-ketoacid dehydrogenase. In addition,<br />

LA enhances glucose metabolism in insulin-resistant rat skeletal muscle,<br />

and showed potent antioxidant activity by chelating transition metal ions and<br />

increasing cytosolic glutathione and vitamin C levels.


Lipoic Acid Blocks Obesity 279<br />

↓ Hypothalamic AMPK<br />

Food<br />

intake<br />

α−lipoic acid<br />

↑ UCP-1 in adipose tissue ↑ AMPK in muscle<br />

Energy<br />

expenditure Body weight<br />

FIGURE 18.3 Mechanism of body weight regulation by LA. LA reduces food intake and<br />

increases energy expenditure by suppressing hypothalamic AMP-activated protein kinase<br />

(AMPK) activity. It increases energy expenditure by increasing uncoupling protein (UCP)-<br />

1 expression in adipose tissue and by activating AMPK in skeletal muscle.<br />

Administration of LA to rodents reduces food intake and body weight as well<br />

as stimulating whole body energy expenditure. Central administration of very<br />

small amounts of LA increased UCP-1 mRNA in BATs, whereas co-administration<br />

of an AMPK activator, 5-aminoimidazole-4-carboxamide ribonucleoside,<br />

prevented the LA-induced enhancement of energy expenditure and UCP-1 expression.<br />

21 These results indicate that food intake, energy expenditure, and UCP-1<br />

are mediated via hypothalamic AMPK inhibition by treatment with LA (Figure<br />

18.3). Finally, in contrast to its effects on the hypothalamus, LA increased glucose<br />

uptake and fatty acid oxidation in skeletal muscle by activating AMPK. 30<br />

The mechanism by which LA decreases hypothalamic AMPK activity is<br />

presently unknown. However, LA has been found to stimulate glucose transport<br />

and ATP synthesis in peripheral tissues, 31,32 and it may decrease hypothalamic<br />

AMPK activity by increasing glucose uptake and/or its metabolism in the hypothalamus.<br />

LA has been shown to activate ACC by decreasing ACC phosphorylation<br />

in hypothalamics. 21 Since activation of ACC can increase intracellular malonyl-CoA<br />

in hypothalamic neurons, malonyl-CoA may be a key downstream<br />

mediator of AMPK activity responsible for the decrease in food intake by LA<br />

administration as shown by Loftus et al. 16 and Ruderman and Prentki. 17


280 Oxidative Stress and Inflammatory Mechanisms<br />

REGULATION OF ADIPOCYTE DIFFERENTIATION<br />

Obesity is closely correlated with the prevalence of diabetes and cardiovascular<br />

disease. Plasma levels of leptin, TNF-α, and non-esterified fatty acids are elevated<br />

in obesity and substantially contribute to the development of insulin resistance. 1<br />

Obesity is caused not only by hypertrophy of adipose tissue, but also by<br />

adipose tissue hyperplasia that triggers the transformation of pre-adipocytes into<br />

adipocytes. 9<br />

Adipocyte differentiation is a complex process that involves coordinated<br />

expression of specific genes and proteins associated with each stage of differentiation.<br />

This process is regulated by several signaling pathways. 33 Insulin, the<br />

major anabolic hormone, promotes in vivo accumulation of adipose tissue. 34<br />

Structurally unrelated inhibitors of PI3-K, LY294002 and wortmannin, have<br />

been shown to block adipocyte differentiation in a time- and dose-dependent<br />

fashion, 35 suggesting that the IR/Akt signaling pathway is important in transducing<br />

the pro-adipogenic effects of insulin. In contrast, mitogen-activated protein<br />

kinases (MAPKs) such as extracellular signal-regulated kinase (ERK) and c-Jun<br />

N-terminal kinase (JNK) suppress the process of adipocyte differentiation. 36,37<br />

TNF-α is known to exert its anti-adipogenic effects, at least in part, through<br />

activation of the ERK pathway. 36 However, p38K is shown to promote adipocyte<br />

differentiation. 38<br />

The signals that regulate adipogenesis either promote or block the cascades<br />

of transcription factors that coordinate the differentiation process. CCAAT element<br />

binding proteins (C/EBPs)-β and -δ and sterol response element binding<br />

protein-1 (ADD1/ SREBP1) are active during the early stages of the differentiation<br />

process and induce the expression and/or activity of peroxisome proliferatoractivated<br />

receptor-γ, (PPAR-γ), a pivotal coordinator of adipocyte differentiation.<br />

PPAR-γ, PPAR-α, and PPAR-δ are important regulators of lipid metabolism.<br />

Although they share significant structural similarities, the biological effects associated<br />

with each isotype are distinct. For example, PPAR-α and PPAR-δ regulate<br />

fatty acid metabolism, while PPAR-γ controls lipid storage and adipogenesis.<br />

Activated PPAR-γ induces exit from the cell cycle and, in cooperation with<br />

C/EBP-α, stimulates the expression of many metabolic genes such as Glut-4,<br />

lipoprotein lipase (LPL), 39 and adipocyte-specific fatty acid binding protein<br />

(aP2), 40 thus constituting a functional lipogenic adipocyte.<br />

JNK and ERK suppress this process by phosphorylating and thereby attenuating<br />

the transcriptional activity of PPAR-γ. 36,37 Besides these integral members<br />

of the adipogenesis program, other transcription factors such as AP-1 41 and cAMP<br />

responsive element binding protein (CREB) 42 are known to promote adipogenesis,<br />

whereas nuclear factor-κB (NF-κB) suppresses adipocyte differentiation. 43<br />

Therefore, the activity and/or expression of these transcription factors are<br />

attractive pharmacological targets for modulating adipocyte tissue formation and<br />

deposition.


Lipoic Acid Blocks Obesity 281<br />

INHIBITION OF ADIPOCYTE DIFFERENTIATION VIA MITOGEN-<br />

ACTIVATED PROTEIN KINASE PATHWAYS<br />

MAPK SIGNALING PATHWAYS MEDIATE ACTIONS OF LA ON<br />

ADIPOGENESIS<br />

Several lines of evidence indicate that pro-adipogenic transcription factors such<br />

as PPAR-γ and members of the C/EBP family can be negatively regulated by<br />

MAPKs. Epidermal growth factor, platelet-derived growth factor, lipoxygenase-<br />

1 metabolites, and prostaglandin F 2α phosphorylate and attenuate transcriptional<br />

activity of PPAR-γ by activating MAPK signaling pathways. 44–46 Similarly, LA<br />

treatment of pre-adipocytes inhibits the insulin- or hormonal cocktail-induced<br />

transcriptional activity of PPAR-γ and C/EBPα, which is accompanied by strong<br />

activation of ERK and JNK.<br />

Inhibitors of ERK and JNK activity abolish the inhibitory effect of LA on<br />

insulin- or hormonal cocktail-induced adipogenesis. On the other hand, LA hardly<br />

stimulates phosphorylation of insulin receptor (IR) or insulin receptor substrate<br />

(IRS)-1 in pre-adipocytes and adipocytes at the early stages of differentiation. In<br />

particular, upon LA treatment, a transient Akt phosphorylation is detected in preadipocytes<br />

although it is not detectable in adipocytes at early stages of differentiation.<br />

In contrast, insulin strongly activates IR and IRS-1 and induces longlasting<br />

Akt activation in pre-adipocytes and in adipocytes at early stages of<br />

differentiation. These findings exclude possible involvement of Akt activation in<br />

LA-induced inhibition of adipogenesis and demonstrate that LA down-regulates<br />

PPAR-γ and C/EBP-α through activation of MAPK signaling pathways such as<br />

ERK and JNK (Figure 18.4).<br />

MODULATION OF AUXILIARY TRANSCRIPTION FACTORS IN ADIPOGENESIS<br />

Transcriptional activities of AP-1 and CREB are increased in fully differentiated<br />

3T3-L1 adipocytes as well as after 2-hr treatment with a hormonal cocktail in<br />

3T3-L1 pre-adipocytes. AP-1 is involved in transcriptional regulation of aP2 and<br />

LPL genes. 47,48 CREB appears to stimulate transcription of several adipocytespecific<br />

genes such as aP2, fatty acid synthetase, and phosphoenolpyruvate carboxykinase.<br />

42 LA, however, strongly down-regulates AP-1 and CREB activities,<br />

whereas it up-regulates NF-κB activities in pre-adipocytes.<br />

Many anti-adipogenic factors such as pro-inflammatory cytokines, 49 TNF-α, 50<br />

and endrin 51 are also known to up-regulate NF-κB activity, whereas pro-adipogenic<br />

factors such as troglitazone display opposite effects in 3T3-L1 cells. 52<br />

Considering that AP-1, 53 CREB, 54 and NF-κB 55 mediate major downstream effects<br />

of MAPK signaling pathways, our findings suggest that activation of the MAPK<br />

signaling pathways by LA leads to differential regulation of these transcription<br />

factors, which eventually results in decreased expression of the adipocyte-specific<br />

genes, and consequently suppresses adipogenesis.


282 Oxidative Stress and Inflammatory Mechanisms<br />

Insulin<br />

IR/PI3-K/Akt<br />

Early gene expression<br />

(c-Myc, c-Fos, c-Jun)<br />

Clonal expansion<br />

FIGURE 18.4 Regulation of adipocyte differentiation by LA. LA activates both the<br />

IR/PI3-K/Akt pathway and the MAPK pathway. Activation of the MAPK pathway mainly<br />

leads to the inhibition of pro-adipogenic transcription factors such as PPAR-γ, C/EBP-α,<br />

AP-1, and CREB, and the activation of anti-adipogenic factor such as NF-κB. The inhibition<br />

of pro-adipogenic transcription factors decreases transcription of several proteins<br />

including LPL and aP2, which are two major markers of adipogenesis. In contrast, activation<br />

of NF-κB may enhance cytokine release and inhibit adipogenesis. LA may also<br />

prevent the adipogenic process by regulating the expression of immediate early genes,<br />

which are involved in the process of clonal expansion.<br />

MAPK SIGNALING PATHWAYS MEDIATE LA ACTIONS ON CELL CYCLE<br />

<strong>AND</strong> CLONAL EXPANSION<br />

AP-1<br />

CREB<br />

α-Lipoic acid<br />

LPL, aP2<br />

MAPK<br />

PPAR-γ<br />

C/EBP-α<br />

NF-κB<br />

Cytokines<br />

(TNF-α, IL-6)<br />

Preadipocytes<br />

Anti-adipogenesis<br />

Dedifferentiation<br />

Adipocytes<br />

In the course of adipogenesis, one of the first events following hormonal induction<br />

is re-entry of growth-arrested pre-adipocytes into the cell cycle. LA has been<br />

demonstrated to inhibit the process of clonal expansion when induced by insulin<br />

or a hormonal cocktail, indicating that insulin and LA oppositely regulate cell<br />

cycle progression. This differential effect seems to be due to the potency and/or<br />

the kinetics of activating of MAPK and IR/Akt signaling pathways.<br />

Both insulin and LA activate MAPK signaling pathways in pre-adipocytes.<br />

However, insulin, but not LA, also strongly activates the IR/Akt signaling pathway.<br />

This observation indicates that progression in the cell cycle and clonal<br />

expansion may require activation of both MAPK and IR/Akt signaling pathways.<br />

On the other hand, insulin-induced MAPK activation is transient, whereas that<br />

of LA lasts longer, indicating that duration of MAPK activation may be another<br />

important factor in determining the fate of a cell in the cell cycle. Indeed, transient<br />

activation of MAPK has been considered as a contributor to cell cycle progression<br />

whereas its prolonged activation can result in cell cycle arrest via induction of<br />

p21 Cip1/Waf1 expression and inhibition of cyclin-dependent kinase activity. 56,57


Lipoic Acid Blocks Obesity 283<br />

It should be emphasized that JNK is known to activate p53, which triggers<br />

activation of several proteins involved in cell cycle arrest such as p21 Cip1/Waf1 . 58<br />

This evidence supports the notion that activation of MAPKs mediates the inhibitory<br />

effect of LA on the clonal expansion process by suppressing the expression<br />

of immediate early genes.<br />

SUMMARY<br />

AMPK is an enzyme that functions as an intracellular energy sensor. It is activated<br />

when the energy status of a cell is low. LA decreases hypothalamic AMPK activity<br />

and results in profound weight losses in rodents by reducing food intake and enhancing<br />

energy expenditures. However, LA increases AMPK activity in skeletal muscle<br />

and other organs, which enhances glucose uptake and mitochondrial fatty acid oxidation.<br />

Activation of hypothalamic AMPK activity reverses by LA administration,<br />

which in turn reduces food intake and increases energy expenditure. ICV administration<br />

of glucose decreases hypothalamic AMPK activity, whereas inhibition of<br />

intracellular glucose utilization through the administration of 2-DG increases hypothalamic<br />

AMPK activity and food intake. 2-DG-induced hyperphagia is reversed by<br />

inhibiting hypothalamic AMPK. These findings indicate that hypothalamic AMPK<br />

is important in the regulation of food intake and energy expenditure and that LA<br />

exerts anti-obesity effects by suppressing hypothalamic AMPK activity.<br />

It was reported that LA inhibited differentiation of pro-adipocytes induced<br />

by a hormone mixture or troglitazone. Northern blot analysis of cells demonstrated<br />

that this inhibition was accompanied by attenuated expression of aP2 and<br />

LPL. Electrophoretic mobility shift assay and western blot analysis of cells<br />

demonstrated that LA modulates transcriptional activity and/or expression of antior<br />

pro-adipogenic transcriptional factors.<br />

LA treatment of pre-adipocytes also resulted in prolonged activation of major<br />

MAPK signaling pathways. These findings suggest that LA inhibits insulin- or<br />

hormonal mixture-induced differentiation of pre-adipocytes by modulating activity<br />

and/or expression of pro- or anti-adipogenic transcriptional factors mainly by<br />

activating the MAPK pathways.<br />

In conclusion, LA may be beneficial in obesity via two major mechanisms.<br />

It may regulate food intake and energy expenditure by decreasing hypothalamic<br />

AMPK activity and block adipocyte differentiation by activating MAPK pathways.<br />

Further studies are warranted to evaluate food intake and energy expenditure<br />

via hypothalamic AMPK activity and adipocyte differentiation by MAPK<br />

pathways in humans.<br />

ACKNOWLEDGMENTS<br />

We thank Drs. Ki-Up Lee and Jong-Yeol Park of the University of Ulsan College<br />

of Medicine, Asan Medical Center, Seoul, Republic of Korea, for providing their<br />

data. This study was supported by Brain Korea 21 Grant M103300.


284 Oxidative Stress and Inflammatory Mechanisms<br />

REFERENCES<br />

1. Leong KS and Wilding JP. Obesity and diabetes. Baillieres Best Pract Res Clin<br />

Endocrinol Metab 1999; 13, 221.<br />

2. Zhang Y, Proenca R, Maffei M, Barone M, Leopold L, and Friedman JM. Positional<br />

cloning of the mouse obese gene and its human homologue. Nature 1994;<br />

372, 425.<br />

3. Tartaglia LA et al. Identification and expression cloning of a leptin receptor, OB-<br />

R. Cell 1995; 83, 1263.<br />

4. Berg A and de Kok A. 2-Oxo acid dehydrogenase multienzyme complexes: the<br />

central role of the lipoyl domain. J Biol Chem 1997; 378, 617.<br />

5. Czech MP and Fain JN. Cu ++ -dependent thiol stimulation of glucose metabolism<br />

in white fat cells. J Biol Chem 1972; 247, 6218.<br />

6. Krieger-Brauer HI, Medda PK, and Kather H. Insulin-induced activation of<br />

NADPH-dependent H 2O 2 generation in human adipocyte plasma membranes is<br />

mediated by G-alpha-i2. J Biol Chem 1997; 272, 10135.<br />

7. Mahadev K et al. Hydrogen peroxide generated during cellular insulin stimulation<br />

is integral to activation of the distal insulin signaling cascade in 3T3-L1 adipocytes.<br />

J Biol Chem 2001; 276, 48662.<br />

8. Lee WJ et al. Obesity: the role of hypothalamic AMP-activated protein kinase in<br />

body weight regulation. Int J Biochem Cell Biol 2005; 37, 2254.<br />

9. Caro JF, Dohm LG, Pories WJ, and Sinha MK. Cellular alterations in liver, skeletal<br />

muscle, and adipose tissue responsible for insulin resistance in obesity and type<br />

II diabetes. Diabetes Metab Rev 1989; 5, 665.<br />

10. Havel PJ. Update on adipocyte hormones: regulation of energy balance and carbohydrate/lipid<br />

metabolism. Diabetes 2004; 53, Suppl 1, S143.<br />

11. Yamauchi T et al. Cloning of adiponectin receptors that mediate antidiabetic<br />

metabolic effects. Nature 2003; 423, 762.<br />

12. Bachman ES et al. Beta AR signaling required for diet-induced thermogenesis<br />

and obesity resistance. Science 2002; 297, 843.<br />

13. Flier JS. Obesity wars: molecular progress confronts an expanding epidemic. Cell<br />

2004; 116, 337.<br />

14. Obici S et al. Central administration of oleic acid inhibits glucose production and<br />

food intake. Diabetes 2002; 51, 271.<br />

15. Sergeyev V, Broberger C, Gorbatyuk O, and Hokfelt T. Effect of 2-mercaptoacetate<br />

and 2-deoxy-D-glucose administration on the expression of NPY, AGRP, POMC,<br />

MCH and hypocretin/orexin in the rat hypothalamus. Neuroreport 2000; 11, 117.<br />

16. Loftus TM et al. Reduced food intake and body weight in mice treated with fatty<br />

acid synthase inhibitors. Science 2000; 288, 2379.<br />

17. Ruderman N and Prentki M. AMP kinase and malonyl-CoA: targets for therapy<br />

of the metabolic syndrome. Nat Rev Drug Discov 2004; 3, 340.<br />

18. Obici S, Feng Z, Arduini A, Conti R, and Rossetti L. Inhibition of hypothalamic<br />

carnitine palmitoyltransferase-1 decreases food intake and glucose production.<br />

Nat Med 2003; 9, 756.<br />

19. Yang XJ, Kow LM, Funabashi T, and Mobbs CV. Hypothalamic glucose sensor:<br />

similarities to and differences from pancreatic beta-cell mechanisms. Diabetes<br />

1999; 48, 1763.


Lipoic Acid Blocks Obesity 285<br />

20. Spanswick D, Smith MA, Mirshamsi S, Routh VH, and Ashford ML. Insulin<br />

activates ATP-sensitive K + channels in hypothalamic neurons of lean, but not obese<br />

rats. Nat Neurosci 2000; 3, 757.<br />

21. Kim MS et al. Anti-obesity effects of alpha-lipoic acid mediated by suppression<br />

of hypothalamic AMP-activated protein kinase. Nat Med 2004; 10, 727.<br />

22. Lee WJ et al. Alpha-lipoic acid prevents endothelial dysfunction in obese rats via<br />

activation of AMP-activated protein kinase. Arterioscler Thromb Vasc Biol 2005;<br />

25, 2488.<br />

23. Dalgaard LT and Pedersen O. Uncoupling proteins: functional characteristics and<br />

role in the pathogenesis of obesity and type II diabetes. Diabetologia 2001; 44,<br />

946.<br />

24. Schwartz MW, Woods SC, Porte D, Jr., Seeley RJ, and Baskin DG. Central nervous<br />

system control of food intake. Nature 2000; 404, 661.<br />

25. Kahn BB, Alquier T, Carling D, and Hardie DG. AMP-activated protein kinase:<br />

ancient energy gauge provides clues to modern understanding of metabolism. Cell<br />

Metab 2005; 1, 15.<br />

26. Kefas BA et al. AICA-riboside induces apoptosis of pancreatic beta cells through<br />

stimulation of AMP-activated protein kinase. Diabetologia 2003; 46, 250.<br />

27. Diraison F et al. Over-expression of sterol-regulatory-element-binding protein-1c<br />

(SREBP1c) in rat pancreatic islets induces lipogenesis and decreases glucosestimulated<br />

insulin release: modulation by 5-aminoimidazole-4-carboxamide ribonucleoside<br />

(AICAR). Biochem J 2004; 378, 769.<br />

28. Andersson U et al. AMP-activated protein kinase plays a role in the control of<br />

food intake. J Biol Chem 2004; 279, 12005.<br />

29. Minokoshi Y et al. AMP-kinase regulates food intake by responding to hormonal<br />

and nutrient signals in the hypothalamus. Nature 2004; 428, 569.<br />

30. Lee WJ et al. Alpha-lipoic acid increases insulin sensitivity by activating AMPK<br />

in skeletal muscle. Biochem Biophys Res Commun 2005; 332, 885.<br />

31. Yaworsky K, Somwar R, Ramlal T, Tritschler HJ, and Klip A. Engagement of the<br />

insulin-sensitive pathway in the stimulation of glucose transport by alpha-lipoic<br />

acid in 3T3-L1 adipocytes. Diabetologia 2000; 43, 294.<br />

32. Cho KJ et al. Alpha-lipoic acid inhibits adipocyte differentiation by regulating<br />

pro-adipogenic transcription factors via mitogen-activated protein kinase pathways.<br />

J Biol Chem 2003; 278, 34823.<br />

33. Torti FM, Torti SV, Larrick JW, and Ringold GM. Modulation of adipocyte<br />

differentiation by tumor necrosis factor and transforming growth factor beta. J<br />

Cell Biol 1989; 108, 1105.<br />

34. Zhang B et al. Insulin- and mitogen-activated protein kinase-mediated phosphorylation<br />

and activation of peroxisome proliferator-activated receptor gamma. J<br />

Biol Chem 1996; 271, 31771.<br />

35. Xia X and Serrero G. Inhibition of adipose differentiation by phosphatidylinositol<br />

3-kinase inhibitors. J Cell Physiol 1999; 178, 9.<br />

36. Font de Mora J, Porras A, Ahn N, and Santos E. Mitogen-activated protein kinase<br />

activation is not necessary for, but antagonizes, 3T3-L1 adipocytic differentiation.<br />

Mol Cell Biol 1997; 17, 6068.<br />

37. Camp HS, Tafuri SR, and Leff T. c-Jun N-terminal kinase phosphorylates peroxisome<br />

proliferator-activated receptor-gamma1 and negatively regulates its transcriptional<br />

activity. Endocrinology 1999; 140, 392.


286 Oxidative Stress and Inflammatory Mechanisms<br />

38. Engelman JA, Lisanti MP, and Scherer PE. Specific inhibitors of p38 mitogenactivated<br />

protein kinase block 3T3-L1 adipogenesis. J Biol Chem 1998; 273,<br />

32111.<br />

39. Schoonjans K et al. PPAR alpha and PPAR gamma activators direct a distinct<br />

tissue-specific transcriptional response via a PPRE in the lipoprotein lipase gene.<br />

Embo J 1996; 15, 5336.<br />

40. Tontonoz P, Hu E, Graves RA, Budavari AI, and Spiegelman BM. mPPAR gamma<br />

2: tissue-specific regulator of an adipocyte enhancer. Genes Dev 1994; 8, 1224.<br />

41. Yang VW, Christy RJ, Cook JS, Kelly TJ, and Lane MD. Mechanism of regulation<br />

of the 422(aP2) gene by cAMP during preadipocyte differentiation. Proc Natl<br />

Acad Sci USA 1989; 86, 3629.<br />

42. Reusch JE, Colton LA, and Klemm DJ. CREB activation induces adipogenesis<br />

in 3T3-L1 cells. Mol Cell Biol 2000; 20, 1008.<br />

43. Ruan H, Hacohen N, Golub TR, Van Parijs L, and Lodish HF. Tumor necrosis<br />

factor-alpha suppresses adipocyte-specific genes and activates expression of preadipocyte<br />

genes in 3T3-L1 adipocytes: nuclear factor-kappaB activation by TNFalpha<br />

is obligatory. Diabetes 2002; 51, 1319.<br />

44. Camp HS and Tafuri SR. Regulation of peroxisome proliferator-activated receptor<br />

gamma activity by mitogen-activated protein kinase. J Biol Chem 1997; 272,<br />

10811.<br />

45. Hsi LC, Wilson L, Nixon J, and Eling TE. 15-lipoxygenase-1 metabolites downregulate<br />

peroxisome proliferator-activated receptor gamma via the MAPK signaling<br />

pathway. J Biol Chem 2001; 276, 34545.<br />

46. Reginato MJ, Krakow SL, Bailey ST, and Lazar MA. Prostaglandins promote and<br />

block adipogenesis through opposing effects on peroxisome proliferator-activated<br />

receptor gamma. J Biol Chem 1998; 273, 1855.<br />

47. Distel RJ, Ro HS, Rosen BS, Groves DL, and Spiegelman BM. Nucleoprotein<br />

complexes that regulate gene expression in adipocyte differentiation: direct participation<br />

of c-fos. Cell 1987; 49, 835.<br />

48. Homma H et al. Estrogen suppresses transcription of lipoprotein lipase gene.<br />

Existence of a unique estrogen response element on the lipoprotein lipase promoter.<br />

J Biol Chem 2000; 275, 11404.<br />

49. Renard P and Raes M. The proinflammatory transcription factor NF-kappa-B: a<br />

potential target for novel therapeutical strategies. Cell Biol Toxicol 1999; 15, 341.<br />

50. Schmid E, Hotz-Wagenblatt A, Hacj V, and Droge W. Phosphorylation of the<br />

insulin receptor kinase by phosphocreatine in combination with hydrogen peroxide:<br />

the structural basis of redox priming. FASEB J 1999; 13, 1491.<br />

51. Mahadev K, Zilbering A, Zhu L, and Goldstein BJ. Insulin-stimulated hydrogen<br />

peroxide reversibly inhibits protein-tyrosine phosphatase 1b in vivo and enhances<br />

the early insulin action cascade. J Biol Chem 2001; 276, 21938.<br />

52. Sen CK, Roy S, Han D, and Packer L. Regulation of cellular thiols in human<br />

lymphocytes by alpha-lipoic acid: a flow cytometric analysis. Free Radic Biol<br />

Med 1997; 22, 1241.<br />

53. Czech MP, Lawrence JC, Jr., and Lynn WS. Evidence for the involvement of<br />

sulfhydryl oxidation in the regulation of fat cell hexose transport by insulin. Proc<br />

Natl Acad Sci USA 1974; 71, 4173.<br />

54. Wilden PA and Pessin JE. Differential sensitivity of the insulin-receptor kinase<br />

to thiol and oxidizing agents in the absence and presence of insulin. Biochem J<br />

1987; 245, 325.


Lipoic Acid Blocks Obesity 287<br />

55. Clark S and Konstantopoulos N. Sulphydryl agents modulate insulin- and epidermal<br />

growth factor (EGF)-receptor kinase via reaction with intracellular receptor<br />

domains: differential effects on basal versus activated receptors. Biochem J 1993;<br />

292, 217.<br />

56. Chen JJ, Kosower NS, Petryshyn R, and London IM. The effects of N-ethylmaleimide<br />

on the phosphorylation and aggregation of insulin receptors in the isolated<br />

plasma membranes of 3T3-F442A adipocytes. J Biol Chem 1986; 261, 902.<br />

57. Jhun BH, Hah JS, and Jung CY. Phenylarsine oxide causes an insulin-dependent,<br />

GLUT4-specific degradation in rat adipocytes. J Biol Chem 1991; 266, 22260.<br />

58. Yang J, Clark AE, Harrison R, Kozka IJ, and Holman GD. Trafficking of glucose<br />

transporters in 3T3-L1 cells. Inhibition of trafficking by phenylarsine oxide implicates<br />

a slow dissociation of transporters from trafficking proteins. Biochem J 1992;<br />

281, 809.


19<br />

CONTENTS<br />

Effects of Conjugated<br />

Linoleic Acid and Lipoic<br />

Acid on Insulin Action in<br />

Insulin-Resistant Obese<br />

Zucker Rats<br />

Erik J. Henriksen<br />

Abstract .............................................................................................................289<br />

Introduction .......................................................................................................290<br />

Regulation of Skeletal Muscle Glucose Transport System..............................291<br />

Etiology of Insulin Resistance in Obese Zucker Rats .....................................292<br />

Nutriceutical Interventions: Conjugated Linoleic Acid and α-Lipoic Acid ....292<br />

Effects of CLA in Obese Zucker Rats....................................................292<br />

Effects of ALA in Obese Zucker Rats....................................................293<br />

Interactions of CLA and ALA in Obese Zucker Rats............................294<br />

Perspectives and Future Directions ..................................................................295<br />

References .........................................................................................................296<br />

ABSTRACT<br />

Insulin-resistant conditions such as pre-diabetes and type 2 diabetes are characterized<br />

by defects in the ability of insulin to activate glucose transport in<br />

skeletal muscle. One animal model that has proven useful in elucidating the<br />

multifactorial etiology of skeletal muscle insulin resistance is the obese Zucker<br />

(fa/fa) rat, characterized by complete leptin resistance, massive central obesity,<br />

hyperinsulinemia, dyslipidemia, and oxidative stress (the imbalance between<br />

exposure of tissue to an oxidant stress and cellular antioxidant defenses).<br />

Studies published by our research group addressed the utility of two nutriceutical<br />

compounds, conjugated linoleic acid (CLA) and alpha-lipoic acid (ALA),<br />

both of which possess antioxidant properties, in improving the metabolic<br />

289


290 Oxidative Stress and Inflammatory Mechanisms<br />

conditions of obese Zucker rats. These studies indicate that chronic administration<br />

of the R-enantiomer of ALA (R-ALA) to obese Zucker rats improved<br />

whole-body insulin sensitivity and enhanced the insulin-stimulated skeletal<br />

muscle glucose transport system, at least in part by up-regulation of IRS-1dependent<br />

insulin signaling. CLA treatment of obese Zucker rats improved<br />

glucose tolerance and insulin-stimulated glucose transport activity, attributable<br />

exclusively to the trans-10, cis-12 isomer and associated with reductions in<br />

oxidative stress and muscle lipid levels. Significant interactions exist between<br />

CLA and R-ALA for enhancement of insulin action on skeletal muscle glucose<br />

transport in the obese Zucker rat, also ascribed to reductions in muscle oxidative<br />

stress and lipid storage. Collectively, these investigations support the<br />

fundamental concept that oxidative stress is an important component of the<br />

etiology of insulin resistance that can be beneficially modulated by antioxidant<br />

interventions, including CLA and ALA.<br />

INTRODUCTION<br />

Insulin resistance of skeletal muscle glucose disposal is a primary defect associated<br />

with the development of the pre-diabetic state and with the conversion from<br />

a pre-diabetic state to overt type 2 diabetes. 1,2 Pre-diabetes is defined as a condition<br />

in which fasting blood glucose levels are elevated above normal (100 to<br />

125 mg/dl) but do not exceed the threshold for the diagnosis of diabetes (>126<br />

mg/dl). 3 Pre-diabetes is frequently characterized by several other cardiovascular<br />

risk factors including visceral obesity, hyperinsulinemia, and dyslipidemia,<br />

encompassing a condition of elevated cardiovascular disease risk known as metabolic<br />

syndrome (see Chapter 1). It is estimated over 40 million individuals in<br />

the United States have pre-diabetes. 3 In the face of further metabolic destabilization<br />

(primarily via ß cell dysfunction), this population could increase substantially<br />

the number of overt type 2 diabetics (presently estimated to be >20 million)<br />

in the U.S. While the etiology of the skeletal muscle insulin resistance present<br />

in pre-diabetic and type 2 diabetic individuals is clearly multifactorial, accumulating<br />

evidence indicates that one factor that can cause insulin resistance and<br />

exacerbate existing insulin resistance is oxidative stress, the imbalance between<br />

exposure of tissues to oxidant stresses and cellular antioxidant defenses. 4<br />

A major challenge to the scientific community and to healthcare professionals<br />

is the design and implementation of effective interventions to reduce insulin<br />

resistance in these pre-diabetic and type 2 diabetic populations. Increased physical<br />

activity and loss of visceral fat have long been the preferred interventions for<br />

ameliorating insulin resistance, 5 but these lifestyle changes are difficult to implement<br />

and maintain in human pre-diabetic and type 2 diabetic subjects. Therefore,<br />

pharmaceutical and nutriceutical treatments are seen as viable alternative therapies<br />

for reducing insulin resistance in these individuals.<br />

Two nutriceutical compounds that received considerable attention recently<br />

as interventions in insulin-resistant states are conjugated linoleic acid (CLA)


Effects of Conjugated Linoleic Acid and Lipoic Acid 291<br />

and alpha-lipoic acid (ALA). 6,7 Both substances have antioxidant properties,<br />

and their utility in the context of oxidative stress-associated insulin resistance<br />

of skeletal muscle glucose disposal will be the focus of this chapter. The normal<br />

regulation of the glucose transport system in skeletal muscle will be described<br />

first, followed by a brief section summarizing the fundamental defects of the<br />

glucose transport system, including the contribution of oxidative stress, that<br />

lead to the insulin-resistant state and ultimately to the development of overt<br />

type 2 diabetes. Finally, a discussion of potential physiological and pharmacological<br />

interventions in the prevention and treatment of insulin resistance is<br />

included, culminating with more detailed coverage of how CLA and ALA,<br />

individually and in combination, can modify oxidative stress and reduce insulin<br />

resistance of skeletal muscle glucose transport in the pre-diabetic state. This<br />

discussion will focus on the obese Zucker (fa/fa) rat, an animal model that<br />

displays complete leptin resistance, massive visceral obesity, glucose intolerance,<br />

insulin resistance, hyperinsulinemia, dyslipidemia, and oxidative stress,<br />

and clearly reflects many of the pathophysiological conditions present in<br />

humans with pre-diabetes and metabolic syndrome.<br />

REGULATION OF SKELETAL MUSCLE GLUCOSE<br />

TRANSPORT SYSTEM<br />

The acute regulation of the insulin-dependent glucose transport system in mammalian<br />

skeletal muscle occurs via the sequential activation of several intracellular<br />

proteins. 8,9 This cascade is initiated by insulin binding to the α subunit of the<br />

insulin receptor (IR), thereby activating the intrinsic tyrosine kinase of the transmembrane<br />

IR ß subunits. The activated IR tyrosine kinase can phosphorylate a<br />

number of important intracellular substrates, including isoforms of insulin receptor<br />

substrate (IRS1-4). The most important isoforms in skeletal muscle are primarily<br />

IRS-1 and IRS-2. Tyrosine-phosphorylated IRS-1 or IRS-2 can then interact<br />

with the p85 regulatory subunit of phosphatidylinositol-3-kinase (PI3-kinase),<br />

increasing the catalytic activity of the p110 subunit of PI3-kinase. Phosphoinositide<br />

moieties produced by PI3-kinase subsequently activate 3-phosphoinositidedependent<br />

kinases (PDKs) that then phosphorylate and activate Akt, a serine/threonine<br />

kinase, and atypical protein kinase C isoforms. The activation of these steps<br />

eventually results in the translocation of the glucose transporter protein isoform<br />

GLUT-4 to the sarcolemmal membrane where glucose transport takes place via<br />

a facilitative diffusion process.<br />

Glucose transport in skeletal muscle can also be activated by contractile<br />

activity. 10 The intracellular mechanisms responsible for this insulin-independent<br />

stimulation of glucose transport include the activation of 5-adenosine-monophosphate-activated<br />

protein kinase (AMP kinase), an enzyme stimulated by a decrease<br />

in cellular energy charge, 11–14 and activation of calcium and calmodulin-dependent<br />

protein kinase. 14,15 Ultimately, these contraction-associated signaling factors stimulate<br />

GLUT-4 translocation 16,17 and enhanced glucose transport activity.


292 Oxidative Stress and Inflammatory Mechanisms<br />

ETIOLOGY OF INSULIN RESISTANCE IN OBESE<br />

ZUCKER RATS<br />

The obese Zucker (fa/fa) rat develops numerous pathophysiological conditions<br />

because of a mutation in the leptin receptor gene. 18–20 This defect results in chronic<br />

hyperphagia and eventually in obesity, especially extensive visceral obesity. Skeletal<br />

muscle insulin resistance in obese Zucker rats is reflected in diminished<br />

insulin-stimulated GLUT-4 protein translocation 21,22 and glucose transport activity.<br />

22–24 The compromised glucose transport system in muscles of obese Zucker<br />

rats likely results from specific defects in the insulin signaling cascade including<br />

reduced IRS-1 protein expression 25–27 and diminished insulin-stimulated IRS-1<br />

tyrosine phosphorylation and PI3-kinase activation. 25,27 Moreover, the obese<br />

Zucker rat is markedly dyslipidemic and displays reduced activity of beneficial<br />

protein kinase C isoforms 28 and overactivation of deleterious protein kinase C<br />

isoforms 29,30 that may be important in the etiology of insulin resistance in this<br />

pre-diabetic animal model.<br />

This abnormal functionality of the insulin signaling cascade in skeletal muscle<br />

is also characteristic of humans with pre-diabetes and overt type 2 diabetes.<br />

Defects in human insulin-resistant skeletal muscle include impaired insulin stimulation<br />

of tyrosine phosphorylation of IR and IRS-1 and of the activities of PI3kinase<br />

and Akt 31–34 and reduced GLUT-4 protein translocation. 35,36<br />

The obese Zucker rat also displays characteristics of oxidative stress that may<br />

be associated with its well established insulin-resistant state. For example, intramuscular<br />

triglycerides and protein carbonyls — markers of oxidative damage 37<br />

— are elevated in the skeletal muscles, cardiac muscles, and livers of these<br />

animals 38–40 and are associated with diminutions of insulin action on the IR/IRS-<br />

1-dependent signaling pathway. 27 We have also demonstrated that in vitro oxidant<br />

stress (~90 µM H 2O 2) can directly impair insulin action on IR/IRS-1-dependent<br />

signaling, glucose transport, and glycogen synthesis in otherwise insulin-sensitive<br />

skeletal muscles from lean Zucker rats (Kim, J.S., Saengsirisuwan, V., and Henriksen,<br />

E.J., manuscript submitted), providing additional support for the concept<br />

that oxidative stress can play a role in the etiology of the insulin-resistant state<br />

in skeletal muscle.<br />

NUTRICEUTICAL INTERVENTIONS: CONJUGATED<br />

LINOLEIC ACID <strong>AND</strong> α-LIPOIC ACID<br />

EFFECTS OF CLA IN OBESE ZUCKER RATS<br />

CLA is a dienoic derivative of linoleic acid that exists primarily as one of two<br />

isomers — the cis-10, trans-12 isomer and the trans-10, cis-12 isomer, with<br />

numerous other minor isoforms possible. CLA possesses significant antioxidant<br />

properties 41 and has been used as an intervention against cancer and heart disease.<br />

42 Chronic CLA administration diminished visceral fat in obese Zucker rats 39<br />

and adult humans. 43 In the Zucker diabetic fatty (ZDF) rat, a model of overt type


Effects of Conjugated Linoleic Acid and Lipoic Acid 293<br />

2 diabetes, the trans-10, cis-12 isomer of CLA enhanced glucose tolerance,<br />

reduced fasting plasma glucose, insulin, and free fatty acids, and increased insulin-stimulated<br />

skeletal muscle glucose transport activity and glycogen synthase<br />

activity. 44,45 This supports its potential as an intervention for treatment of type 2<br />

diabetes, possibly as an activator of the peroxisome proliferator-activated receptor-γ<br />

in a variety of tissues. 44<br />

Our research group conducted investigations of the metabolic actions of CLA<br />

in pre-diabetic obese Zucker rats. In this model of pre-diabetes, as in diabetic<br />

ZDF rats, 44,45 CLA treatment led to enhanced glucose tolerance, increased wholebody<br />

insulin sensitivity, decreased fasting plasma glucose, insulin, and free fatty<br />

acid levels, and improved insulin-mediated skeletal muscle glucose transport<br />

activity. 39,40 Importantly, CLA treatment caused a significant decrease in visceral<br />

fat in this grossly obese animal model. 39<br />

These metabolic actions of CLA can be attributed solely to the trans-10, cis-<br />

12 isomer, as the cis-10, trans-12 isomer is metabolically neutral in this animal<br />

model. Moreover, these trans-10, cis-12-CLA-induced metabolic improvements<br />

in the obese Zucker rat are highly correlated with reductions in skeletal muscle<br />

protein carbonyl levels and intramuscular triglyceride levels. 39 This latter finding<br />

is of importance in light of the well-recognized role of elevated muscle lipid in<br />

the etiology of insulin resistance. 46 Collectively, these results highlight the potential<br />

of the trans-10, cis-12 isomer of CLA as a nutriceutical intervention in specific<br />

conditions of glucose intolerance and insulin resistance, and indicate that the<br />

beneficial metabolic actions of CLA may be related to its ability to reduce<br />

intramuscular lipid levels and local indices of oxidative stress.<br />

EFFECTS OF ALA IN OBESE ZUCKER RATS<br />

The metabolic actions of ALA in pre-diabetic obese Zucker rats have been<br />

extensively investigated by our research group in the last decade, and most of<br />

these findings are summarized in recent review articles and chapters. 4,6,47 The<br />

highlights of these studies will be briefly covered in this section.<br />

The vast majority of the beneficial metabolic actions of ALA in obese Zucker<br />

rats can be attributed to the R-enantiomer, with the S-enantiomer being either<br />

metabolically neutral or even eliciting some metabolic worsening such as<br />

decreased muscle GLUT4 glucose transporter protein expression. 48 R-ALA<br />

acutely enhances glucose transport and intracellular glucose metabolism in both<br />

insulin-sensitive 49,50 and insulin-resistant 48,50,51 skeletal muscle preparations.<br />

Chronic administration (2 to 6 weeks) of R-ALA to obese Zucker rats resulted<br />

in improvements of whole-body glucose tolerance and insulin sensitivity, reductions<br />

in fasting plasma insulin and lipids, and enhanced insulin-stimulated glucose<br />

transport activity in isolated skeletal muscle. 27,38<br />

The increased insulin action on skeletal muscle glucose transport associated<br />

with chronic R-ALA treatment of obese Zucker rats is correlated with a proportional<br />

reduction of oxidative stress, as reflected by muscle protein carbonyl<br />

levels. 38 This antioxidant intervention is associated with reductions in cardiac and


294 Oxidative Stress and Inflammatory Mechanisms<br />

liver protein carbonyl levels. 38 Chronic treatment of obese Zucker rats with R-<br />

ALA also induced dose-dependent reductions in plasma free fatty acids that<br />

correlated with improvements in whole-body glucose tolerance and insulin sensitivity<br />

and with improved insulin action on the skeletal muscle glucose transport<br />

system. 38,52 It now appears that R-ALA treatment can improve functionality of<br />

the IRS-1-dependent insulin signaling pathway in skeletal muscles of obese<br />

Zucker rats. Chronic administration of R-ALA to this insulin-resistant, pre-diabetic<br />

animal model induces an enhancement of IRS-1 protein expression and<br />

increased insulin-stimulated IRS-1 associated with the p85 subunit of phosphatidylinositol-3-kinase,<br />

a surrogate measure of PI3-kinase activation. 27<br />

These results on the beneficial effects of chronic treatment with ALA in obese<br />

Zucker rats were confirmed recently in another rodent model of obesity-associated<br />

insulin resistance and glucose dysregulation, the Otsuka Long-Evans Tokushima<br />

Fatty (OLETF) rat. 53 OLETF rats treated with ALA displayed reduced body<br />

weights, decreased blood glucose levels, and lower intramuscular triglyceride<br />

concentrations. 53 These investigators made the novel finding that the reduction in<br />

muscle lipids and the increase in muscle glucose uptake in short-term ALA-treated<br />

OLETF rats may be associated with the activation of AMP kinase. 54<br />

Several clinical investigations generally supported the animal model-based<br />

findings on the effects of ALA on glucoregulation. The acute infusion of ALA<br />

in type 2 diabetic human subjects resulted in ~30% improvement in glucose<br />

disposal rate during a euglycemic, hyperinsulinemic clamp. 55,56 Moreover, the<br />

chronic intravenous or oral administration of ALA to type 2 diabetic subjects<br />

also elicited significant improvements in whole-body insulin sensitivity. 57–59 The<br />

results from this limited number of clinical investigations, combined with the<br />

findings from animal models, provide further support for additional clinical studies<br />

on the effectiveness of ALA in the treatment of insulin resistance in prediabetic<br />

and type 2 diabetic humans.<br />

INTERACTIONS OF CLA <strong>AND</strong> ALA IN OBESE ZUCKER RATS<br />

To our knowledge, there is only one published investigation related to potential<br />

interactive effects of CLA and ALA to modulate insulin action in insulin-resistant<br />

skeletal muscle. The study, performed by our research group, was published in<br />

2003. 40 Obese Zucker rats were treated chronically with either a mixture of CLA<br />

isomers, with R-ALA, or with a combination of R-ALA and CLA, at low,<br />

minimally effective doses, or at high, maximally effective doses.<br />

While the high doses of CLA and R-ALA individually induced significant<br />

metabolic improvements in these pre-diabetic animals, including increased insulin-stimulated<br />

glucose transport activity and reduced muscle oxidative stress and<br />

lipid levels in skeletal muscle, the combination of these high doses of CLA and<br />

R-ALA did not provide for any further metabolic enhancements above those<br />

brought about by the individual agents. In contrast, the combination of low doses<br />

of CLA and R-ALA (which individually produced minimal metabolic actions)<br />

resulted in substantial improvements in glucose tolerance and whole-body insulin


Effects of Conjugated Linoleic Acid and Lipoic Acid 295<br />

sensitivity and caused significant increases in insulin-stimulated glucose transport<br />

activity in skeletal muscle that were closely associated with reductions in muscle<br />

oxidative stress and lipid levels. This novel synergistic interaction between the<br />

fatty acid CLA and the metabolic antioxidant R-ALA supports the use of combined<br />

CLA and R-ALA in the modulation of whole-body and skeletal muscle<br />

insulin resistance.<br />

PERSPECTIVES <strong>AND</strong> FUTURE DIRECTIONS<br />

The information discussed in this chapter supports the beneficial role of the R-<br />

ALA and CLA nutriceutical compounds in the modulation of insulin action in<br />

insulin-resistant, pre-diabetic obese Zucker rats. With regard to these effects of<br />

R-ALA, the general consensus is that this antioxidant and reactive sulfhydrylcontaining<br />

agent elicits improvements in glucose tolerance, insulin sensitivity,<br />

IR/IRS-1-dependent insulin signaling, and muscle glucose transport capacity.<br />

While some variability in the absolute effect of ALA administration on these<br />

variables in both animal models and in human type 2 diabetics does exist in the<br />

literature, this can generally be attributed to differences among studies in the<br />

ALA dose given, the route of ALA administration, the release rate of the ALA,<br />

and the initial degree of insulin resistance present in the experimental subjects.<br />

On the other hand, several investigations utilizing both animal models and<br />

human subjects presented contradictory results arising from chronic CLA administration.<br />

In some investigations, CLA administration actually worsened insulin<br />

sensitivity in obese mice 60,61 and in abdominally obese men. 62,63 At least part of<br />

this controversy can be accounted for by differential responses to the CLA in the<br />

inherently different animal models employed. For example, the chronic administration<br />

of trans-10, cis-12-CLA to obese Zucker rats 39 or ZDF rats 45 caused a<br />

relatively small decrease in fat mass and an increase in whole-body and skeletal<br />

muscle insulin sensitivity, whereas this same CLA administration to obese<br />

C57BL/6 mice elicited a relatively large fat mass loss but induced insulin resistance.<br />

60,61 The fundamentally different metabolic response to CLA in the mouse<br />

model can likely be ascribed to the development of a state of lipodystrophy<br />

marked by an extreme fat mass deficit and impaired insulin action. In contrast,<br />

the CLA-treated obese Zucker and ZDF rats retained a morbidly obese phenotype.<br />

The specific animal model employed and human population under investigation<br />

must be taken into consideration when assessing the potential for CLA to bring<br />

about beneficial modulation of metabolic parameters.<br />

Finally, based on the results of our study in obese Zucker rats, 40 an important<br />

area for future clinical investigation would be the assessment of the effects of<br />

combined low doses of CLA and R-ALA on metabolic parameters in obese<br />

humans with insulin resistance, glucose intolerance, and dyslipidemia. There are<br />

novel interactions of these two nutriceutical compounds that may be exploited in<br />

the treatment of human insulin-resistant states, especially those states associated<br />

with visceral adiposity.


296 Oxidative Stress and Inflammatory Mechanisms<br />

REFERENCES<br />

1. Warram, J.H. et al. Slow glucose removal rate and hyperinsulinemia precede the<br />

development of type II diabetes in the offspring of diabetic parents. Ann. Intern.<br />

Med. 113, 909, 1990.<br />

2. Ferrannini, E. Insulin resistance versus insulin deficiency in non-insulin-dependent<br />

diabetes mellitus: problems and prospects. Endocr. Rev. 19, 477, 1998.<br />

3. American Diabetes Association. Diagnosis and classification of diabetes mellitus.<br />

Diabetes Care 27, S5, 2004.<br />

4. Henriksen, E.J. Oxidative stress and antioxidant treatment: effects on muscle<br />

glucose transport in animal models of type 1 and type 2 diabetes, in Antioxidants<br />

in Diabetes Management, Packer, L. et al., Eds., Marcel Dekker, New York, 2000.<br />

5. Henriksen, E.J. Invited review: effects of acute exercise and exercise training of<br />

insulin resistance. J. Appl. Physiol. 93, 788, 2002.<br />

6. Henriksen, E.J. Therapeutic effects of lipoic acid on hyperglycemia and insulin<br />

resistance, in Handbook of Antioxidants, Cadenas, E. and Packer, L., Eds., Marcel<br />

Dekker, New York, 2001.<br />

7. Belury, M.A. Dietary conjugated linoleic acid in health: physiological effects and<br />

mechanisms of action. Ann. Rev. Nutr. 22, 505, 2002.<br />

8. Shepherd, P.R. and Kahn, B.B. Glucose transporters and insulin action: implications<br />

for insulin resistance and diabetes mellitus. New Engl. J. Med. 341, 248,<br />

1999.<br />

9. Zierath, J.R., Krook, A., and Wallberg-Henriksson, H. Insulin action and insulin<br />

resistance in human skeletal muscle. Diabetologia 43, 821, 2000.<br />

10. Jessen, N. and Goodyear, L.J. Contraction signaling to glucose transport in skeletal<br />

muscle. J. Appl. Physiol. 99, 330, 2005.<br />

11. Kurth-Kraczek, E.J. et al. 5 AMP-activated protein kinase activation causes<br />

GLUT4 translocation in skeletal muscle. Diabetes 48, 1667, 1999.<br />

12. Mu, J. et al. A role for AMP-activated protein kinase in contraction- and hypoxiaregulated<br />

glucose transport in skeletal muscle. Molecular Cell 7, 1085, 2001.<br />

13. Winder, W.W. Energy-sensing and signaling by AMP-activated protein kinase in<br />

skeletal muscle. J. Appl. Physiol. 91, 1017, 2001.<br />

14. Wright, D.C. et al. Ca 2+ and AMPK both mediate stimulation of glucose transport<br />

by muscle contractions. Diabetes 53, 330, 2004.<br />

15. Holloszy J.O. and Hansen, P.A. Regulation of glucose transport into skeletal<br />

muscle. Rev. Physiol. Biochem. Pharmacol. 128, 99, 1996.<br />

16. Goodyear, L.J., Hirshman, M.F., and Horton, E.S. Exercise-induced translocation<br />

of skeletal muscle glucose transporters. Am. J. Physiol. Endocrinol. Metab. 261,<br />

E795, 1991.<br />

17. Gao, J. et al. Additive effect of contractions and insulin on GLUT-4 translocation<br />

into the sarcolemma. J. Appl. Physiol. 77, 1587, 1994.<br />

18. Phillips, M.S. et al. Leptin receptor missense mutation in the fatty Zucker rat.<br />

Nat. Genetics 13, 18, 1996.<br />

19. Iida, M. et al. Substitution at codon 269 (glutamine proline) of the leptin receptor<br />

(OB-R) cDNA is the only mutation found in the Zucker fatty (fa/fa) rat. Biochem.<br />

Biophys. Res. Commun. 224, 597, 1996.<br />

20. Takaya, K. et al. Molecular cloning of rat leptin receptor isoform complementary<br />

DNAs: identification of a missense mutation in Zucker fatty (fa/fa) rats. Biochem.<br />

Biophys. Res. Commun. 225, 75, 1996.


Effects of Conjugated Linoleic Acid and Lipoic Acid 297<br />

21. King, P.A. et al. Exercise, unlike insulin, promotes glucose transporter translocation<br />

in obese Zucker rat muscle. Am. J. Physiol. Regulatory Integrative Comp.<br />

Physiol. 265, R447, 1993.<br />

22. Etgen, G.J. et al. Glucose transport and cell surface GLUT-4 protein in skeletal<br />

muscle of the obese Zucker rat. Am. J. Physiol. Endocrinol. Metab. 271, E294,<br />

1996.<br />

23. Crettaz, M. et al. Insulin resistance in soleus muscle from obese Zucker rats:<br />

involvement of several defective sites. Biochem. J. 186, 525, 1980.<br />

24. Henriksen, E.J. and Jacob, S. Effects of captopril on glucose transport activity in<br />

skeletal muscle of obese Zucker rats. Metabolism 44, 267, 1995.<br />

25. Anai, M. et al. Altered expression levels and impaired steps in pathways to<br />

phosphatidylinositol-3-kinase activation via insulin receptor substrates 1 and 2 in<br />

Zucker fatty rats. Diabetes 47, 13, 1998.<br />

26. Hevener, A.L., Reichart, D., and Olefsky, J., Exercise and thiazolidinedione therapy<br />

normalize insulin action in the obese Zucker fatty rat. Diabetes 49, 2154, 2000.<br />

27. Saengsirisuwan, V. et al. Interactions of exercise training and R-(+)-alpha-lipoic<br />

acid on insulin signaling in skeletal muscle of obese Zucker rats. Am. J. Physiol.<br />

Endocrinol. Metab. 287, E529, 2004.<br />

28. Cooper, D.R., Watson, J.E., and Dao, M.L. Decreased expression of protein kinase-<br />

C alpha, beta, and epsilon in soleus muscle of Zucker obese (fa/fa) rats. Endocrinology<br />

133, 2241, 1993.<br />

29. Avignon, A. et al. Chronic activation of protein kinase C in soleus muscles and<br />

other tissues of insulin-resistant type II diabetic Goto-Kakizaki (GK), obese/aged,<br />

and obese/Zucker rats: a mechanism for inhibiting glycogen synthesis. Diabetes<br />

45, 1396, 1996.<br />

30. Qu, X., Seale, J.P., and Donnelly, R. Tissue and isoform-selective activation of<br />

protein kinase C in insulin-resistant obese Zucker rats: effects of feeding. J.<br />

Endocrinol. 162, 207, 1999.<br />

31. Goodyear, L.J. et al. Insulin receptor phosphorylation, insulin receptor substrate-<br />

1 phosphorylation, and phosphatidylinositol 3-kinase activity are decreased in<br />

intact skeletal muscle strips from obese subjects. J. Clin. Invest. 95, 2195, 1995.<br />

32. Björnholm, M. et al. Insulin receptor substrate-1 phosphorylation and phosphatidylinositol-3-kinase<br />

activity in skeletal muscle from NIDDM subjects after in vivo<br />

insulin stimulation. Diabetes 46, 524, 1997.<br />

33. Krook, A. et al. Insulin-stimulated Akt kinase activity is reduced in skeletal muscle<br />

from NIDDM subjects. Diabetes 47, 1281, 1998.<br />

34. Krook, A. et al. Characterization of signal transduction and glucose transport in<br />

skeletal muscle from type 2 diabetic patients. Diabetes 49, 284, 2000.<br />

35. Zierath, J.R. et al. Insulin action on glucose transport and plasma membrane<br />

GLUT4 content in skeletal muscle from patients with NIDDM. Diabetologia 39,<br />

1180, 1996.<br />

36. Ryder, J.W. et al. Use of a novel impermeable biotinylated photolabeling reagent<br />

to assess insulin- and hypoxia-stimulated cell surface GLUT4 content in skeletal<br />

muscle from type 2 diabetic patients. Diabetes 49, 647, 2000.<br />

37. Reznick, A.Z. and Packer, L. Oxidative damage to proteins: spectrophotometric<br />

method for carbonyl assay. Methods Enzymol. 233, 357, 1994.<br />

38. Saengsirisuwan, V. et al. Interactions of exercise training and alpha-lipoic acid on<br />

glucose transport in obese Zucker rat. J. Appl. Physiol. 91, 145, 2001.


298 Oxidative Stress and Inflammatory Mechanisms<br />

39. Henriksen, E.J. et al. Isomer-specific actions of conjugated linoleic acid on muscle<br />

glucose transport in the obese Zucker rat. Am. J. Physiol. Endocrinol. Metab. 285,<br />

E98, 2003.<br />

40. Teachey, M.K. et al. Interactions of conjugated linoleic acid and alpha-lipoic acid<br />

on insulin action in the obese Zucker rat. Metabolism 52, 1167, 2003.<br />

41. Leung, Y.H. and Liu, R.H. Trans-10, cis-12-conjugated linoleic acid isomer exhibits<br />

stronger oxyradical scavenging capacity than cis-9, trans-11-conjugated linoleic<br />

acid isomer. J. Agricult. Food Chem. 48, 5469, 2000.<br />

42. Moya-Camarena, S.Y. and Belury, M.A. Species differences in the metabolism<br />

and regulation of gene expression by conjugated linoleic acid. Nutr. Rev. 57, 336,<br />

1999.<br />

43. Blankson, H. et al. Conjugated linoleic acid reduces body fat mass in overweight<br />

and obese humans. J. Nutr. 130, 2943, 2000.<br />

44. Houseknecht, K.L. et al. Dietary conjugated linoleic acid normalizes impaired<br />

glucose tolerance in the Zucker diabetic fatty fa/fa rat. Biochem. Biophys. Res.<br />

Commun. 244, 678, 1998.<br />

45. Ryder, J.W. et al. Isomer-specific antidiabetic properties of conjugated linoleic<br />

acid. Improved glucose tolerance, skeletal muscle insulin action, and UCP-2 gene<br />

expression. Diabetes 50, 1149, 2001.<br />

46. Shulman, G.I. Cellular mechanisms of insulin resistance. J. Clin. Invest. 106, 171,<br />

2000.<br />

47. Henriksen, E.J. Exercise training and the antioxidant alpha-lipoic acid in the<br />

treatment of insulin resistance and type 2 diabetes. Free Rad. Biol. Med. 40, 3,<br />

2006.<br />

48. Streeper, R.S. et al. Differential effects of lipoic acid stereoisomers on glucose<br />

metabolism in insulin-resistant skeletal muscle. Am. J. Physiol. Endocrinol. Metab.<br />

273, E185, 1997.<br />

49. Haugaard, N. and Haugaard, E.S. Stimulation of glucose utilization by thioctic<br />

acid in rat diaphragm incubated in vitro. Biochim. Biophys. Acta 222, 583, 1970.<br />

50. Henriksen, E.J. et al. Stimulation by α-lipoic acid of glucose transport activity in<br />

skeletal muscle of lean and obese Zucker rats. Life Sci. 61, 805, 1997.<br />

51. Jacob, S. et al. The antioxidant α-lipoic acid enhances insulin stimulated glucose<br />

metabolism in insulin-resistant rat skeletal muscle. Diabetes 45, 1024, 1996.<br />

52. Peth, J.A. et al. Effects of a unique conjugate of alpha-lipoic acid and gammalinolenic<br />

acid on insulin action in the obese Zucker rat. Am. J. Physiol. Regulatory<br />

Integrative Comp. Physiol. 278, R453, 2000.<br />

53. Song, K.H. et al. Alpha-lipoic acid prevents diabetes mellitus in diabetes-prone<br />

obese rats. Biochem. Biophys. Res. Commun. 326, 197, 2005.<br />

54. Lee, W.J. et al. Alpha-lipoic acid increases insulin sensitivity by activating AMPK<br />

in skeletal muscle. Biochem. Biophys. Res. Commun. 332, 885, 2005.<br />

55. Rett, K. et al. Effect of acute infusion of thioctic acid on oxidative and nonoxidative<br />

metabolism in obese subjects with NIDDM (abstract). Diabetologia 38,<br />

A41, 1995.<br />

56. Jacob, S. et al. Enhancement of glucose disposal in patients with type 2 diabetes<br />

by alpha-lipoic acid. Drug Res. 45: 872, 1995.<br />

57. Jacob, S. et al. Improvement of insulin-stimulated glucose disposal in type 2<br />

diabetes after repeated parenteral administration of thioctic acid. Exp. Clin. Endocrinol.<br />

Diab. 104, 284, 1996.


Effects of Conjugated Linoleic Acid and Lipoic Acid 299<br />

58. Jacob, S. et al. Oral administration of rac-α-lipoic acid modulates insulin sensitivity<br />

in patients with type 2 diabetes mellitus: a placebo controlled pilot trial.<br />

Free Rad. Biol. Med. 27, 309, 1999.<br />

59. Konrad, T. et al. Alpha-lipoic acid treatment decreases serum lactate and pyruvate<br />

concentrations and improves glucose effectiveness in lean and obese patients with<br />

type 2 diabetes. Diabetes Care 22, 280, 1999.<br />

60. Tsuboyama-Kasaoka, N. et al. Conjugated linoleic acid supplementation reduces<br />

adipose tissue by apoptosis and develops lipodystrophy in mice. Diabetes 49,<br />

1534, 2000.<br />

61. Roche, H.M. et al. Isomer-dependent metabolic effects of conjugated linoleic acid:<br />

insights from molecular markers sterol regulatory element-binding protein-1c and<br />

LXRα. Diabetes 51, 2037, 2002.<br />

62. Riserus, U. et al. Treatment with dietary trans-10, cis-12 conjugated linoleic acid<br />

causes isomer-specific insulin resistance in obese men with the metabolic syndrome.<br />

Diabetes Care 25, 1516, 2002.<br />

63. Riserus, U. et al. Supplementation with conjugated linoleic acid causes isomerdependent<br />

oxidative stress and elevated C-reactive protein: a potential link to fatty<br />

acid-induced insulin resistance. Circulation 106, 1925, 2002.


20<br />

CONTENTS<br />

Trivalent Chromium<br />

Supplementation<br />

Inhibits Oxidative Stress,<br />

Protein Glycosylation,<br />

and Vascular<br />

Inflammation in High<br />

Glucose-Exposed<br />

Human Erythrocytes and<br />

Monocytes<br />

Sushil K. Jain<br />

Abstract .............................................................................................................301<br />

Introduction .......................................................................................................302<br />

Pro-Inflammatory Cytokines and Vascular Inflammation in Diabetes.............302<br />

Oxidative Stress and Vascular Inflammation....................................................303<br />

Trivalent Chromium and Diabetes....................................................................303<br />

Conclusion.........................................................................................................307<br />

Acknowledgments .............................................................................................308<br />

References .........................................................................................................308<br />

ABSTRACT<br />

Epidemiological studies have shown lower levels of chromium among men with<br />

diabetes and cardiovascular disease (CVD) compared with healthy control<br />

301


302 Oxidative Stress and Inflammatory Mechanisms<br />

subjects. The mechanism by which chromium may decrease the incidence of<br />

CVD and insulin resistance is not known. This study demonstrates that chromium<br />

inhibits glycosylation of proteins, oxidative stress, and pro-inflammatory cytokine<br />

secretion, all of which are risk factors in the development of CVD. Erythrocytes<br />

or monocytes isolated from fresh human blood were treated with high concentrations<br />

of glucose (mimicking diabetes) in the presence or absence of chromium<br />

chloride in a medium at 37˚C for 24 hr. We observed that chromium supplementation<br />

prevented increases in protein glycosylation and oxidative stress caused by<br />

high levels of glucose in erythrocytes. In monocytes, chromium supplementation<br />

inhibited the high glucose-induced secretion of interleukin (IL)-6 and tumor<br />

necrosis factor (TNF)-α. This study demonstrates that chromium supplementation<br />

protects against oxidative stress and vascular inflammation caused by exposure<br />

to high glucose in blood cells and provides evidence for a novel mechanism by<br />

which chromium supplementation may decrease the incidence of CVD in diabetic<br />

patients.<br />

INTRODUCTION<br />

Vascular inflammation and CVD are the leading causes of morbidity and mortality<br />

in the diabetic population and remain major public health issues. Hyperglycemia<br />

is one of the major risk factors in the development of vascular complications. 1<br />

Intensive blood glucose control dramatically reduces the devastating complications<br />

that result from poorly controlled diabetes. Diabetes is treated with diet and<br />

insulin administration. However, for many patients, achieving tight glucose control<br />

is difficult with current regimens. The risk of CVD in diabetics is three- to<br />

five-fold greater than that of the general population.<br />

PRO-INFLAMMATORY CYTOKINES <strong>AND</strong> VASCULAR<br />

INFLAMMATION IN DIABETES<br />

TNF-α and IL-6 are pro-inflammatory cytokines produced by macrophages and<br />

other cell types in response to various stimuli. 2,3 The levels of these cytokines<br />

are elevated in the blood of many diabetic patients. 4–10 Increases in circulating<br />

levels of TNF-α and IL-6 decrease insulin sensitivity, 2,10–16 which can necessitate<br />

higher doses of insulin to maintain glycemic control in type 1 diabetic patients.<br />

Increased levels of insulin administration carry a risk of hypoglycemia. In addition,<br />

elevated circulating levels of TNF-α and IL-6 can induce expression of<br />

adhesion molecules and thus monocyte–endothelial cell adhesion, now recognized<br />

as an early and rate-limiting step in vascular inflammation and the development<br />

of vascular disease. 17–23<br />

Different human endothelial cells may have distinct ICAM forms and expression<br />

mechanisms. 24 Expression of VCAM-1 independent of systemic levels of<br />

TNF-α has been shown in pulmonary endothelial cells. 17 Several genes associated<br />

with adhesion molecules (ICAM-1, VCAM-1) and inflammatory cytokines are


Trivalent Chromium Supplementation 303<br />

regulated by NFκB. 25,26 The agents that suppress NFκB activation can diminish<br />

cell adhesion and vascular inflammation.<br />

<strong>OXIDATIVE</strong> <strong>STRESS</strong> <strong>AND</strong> VASCULAR INFLAMMATION<br />

Oxidative stress can also influence the expression of multiple genes in vascular<br />

cells, including signaling molecules such as PKC, NFκB, and ERK. 27–32 Overexpression<br />

of these genes may lead to endothelial dysfunction and ultimately to<br />

microvascular and macrovascular disease. High glucose (HG) can up-regulate<br />

expression of transcription factors, such as NFκB, activating protein-1 and the TNFα<br />

genes in monocytes. 31 TNF-α accumulation in conditioned media increased 10fold<br />

and mRNA levels were increased 11.5-fold by HG. 31 This indicates that both<br />

NFκB and AP-1 mediated enhanced TNF-α transcription by HG. 31<br />

This expression of the TNF-α gene is mediated by the protein kinases p38<br />

and JNK-1, which are respectively dependent on and independent of oxidative<br />

stress pathways. 30,31 Several studies advocate the importance of the p38 pathway<br />

in diabetes. 30 cAMP-dependent protein kinases (PKAs) can activate phosphorylation<br />

of substrate proteins and cross-talk with MAPK pathways and proteins<br />

involved in signal transduction pathways, leading to altered gene expression and<br />

modulation of physiological processes. 32–37 cAMP is known to modulate cytokine<br />

production in a number of cell types. 32–37 This suggests that oxidative stress plays<br />

a key role in the regulatory pathway that progresses from elevated glucose to<br />

monocyte and endothelial cell activation in the enhanced vascular inflammation<br />

of diabetes. 21,38<br />

TRIVALENT CHROMIUM <strong>AND</strong> DIABETES<br />

Trivalent chromium, the reduced form of the element, is required for insulin<br />

action. 39–44 A chromium-containing oligopeptide present in insulin-sensitive cells<br />

binds to the insulin receptor, markedly increasing the activity of the insulinstimulated<br />

tyrosine kinase. 39,40 Overt chromium deficiency has been demonstrated<br />

in patients receiving total parenteral nutrition without chromium supplementation.<br />

45 It is characterized by hyperglycemia, glycosuria, and weight loss that<br />

cannot be controlled with insulin. 45,46 As a consequence, total parenteral nutrition<br />

solutions are regularly supplemented with chromium. 47 Intraperitoneal injections<br />

of potassium chromate also reversed atherosclerotic plaques in rabbits. 48,49<br />

The main route of exposure to chromium in the general population is dietary<br />

intake. Most chromium in the diet is trivalent, and any hexavalent chromium in<br />

food or water is reduced to the trivalent form in the acidic environment of the<br />

stomach. 50,51 Foods with high chromium concentrations include whole grain products,<br />

green beans, broccoli, and bran cereals. 52 The chromium contents of meats,<br />

poultry, and fish vary widely because chromium may be introduced during transport,<br />

processing, and fortification. 52 Foods rich in refined sugars are low in<br />

chromium and actually promote chromium loss. 53


304 Oxidative Stress and Inflammatory Mechanisms<br />

Based on the chromium contents of well balanced diets, 52 adequate intake<br />

values in adults have been established at 35 μg/day for men and 25 μg/day for<br />

women. 51 Although there are no national survey data covering chromium intake, 51<br />

a study of self-selected diets of U.S. adults indicated that the chromium intakes<br />

of a substantial proportion of subjects may be well below the adequate intake. 54<br />

Similar results have been shown in the United Kingdom, Finland, Canada, and<br />

New Zealand. 55 Thus, subclinical chromium deficiency may be a contributor to<br />

glucose intolerance, insulin resistance, and cardiovascular disease, particularly in<br />

aging populations and populations that have increased chromium requirements<br />

because of high sugar diets. 43<br />

Concentrations of chromium in the blood, lenses, and toenails are lower in<br />

diabetic patients compared with concentrations in the normal population. 46,56,57<br />

Several studies have suggested that chromium supplementation may be beneficial<br />

in individuals with type 2 and type 1 diabetes, gestational diabetes, or steroidinduced<br />

diabetes, as evidenced by decreased blood glucose values or decreased<br />

insulin requirements. 42,43,58–67 Although the majority of research on diabetes has<br />

focused on type 2, a few small studies have tested the efficacy of Cr 3+ on type 1<br />

diabetes and found it effective. 64 One study supplemented 162 patients (48 with<br />

type 1 diabetes, the remainder with type 2) with 200 μg/day Cr 3+ picolinate.<br />

Seventy-one percent of the type 1 patients responded positively, allowing 30%<br />

decreases in insulin doses. Blood sugar fluctuations also responded positively,<br />

decreasing as early as 10 days after treatment. Supplementation of chromium as<br />

the picolinate (600 μg/day) in a 28-year-old woman with an 18-year history of<br />

type 1 diabetes reduced HbA 1C from 11.3 to 7.9% in 3 months. 65 When patients<br />

receiving total parenteral therapy were supplemented with chromium, their diabetes<br />

symptoms reversed and they required smaller doses of exogenous insulin. 66<br />

In atherosclerotic rabbits, an injection of chromium chloride resulted in<br />

marked reductions in the plaques covering the aortic intimal surfaces, aortic<br />

weights, and cholesterol contents. 49 Chromium can reduce elevated cholesterol<br />

and triglycerides in a dose-dependent relationship. 68 Results from two casecontrol<br />

studies suggest an inverse association between chromium levels in toenails<br />

and risks of myocardial infarction in the general population. 69 Similarly, a recent<br />

report of the Health Professionals Follow-up Study found lower levels of toenail<br />

chromium among men with diabetes and CVD compared with healthy control<br />

subjects. 56<br />

However, randomized trials of chromium supplementation in diabetes have<br />

not been definitive. Many studies have not been blinded, used inappropriate<br />

glucose metabolism assessment parameters, or included heterogeneous and<br />

poorly characterized patient populations. More rigorous blinded and well controlled<br />

studies are needed to fully assess the efficacy and mechanism of the action<br />

of Cr 3+ supplementation as an adjuvant therapy for diabetic patients.<br />

Figure 20.1 illustrates the effects of high glucose (HG) and trivalent<br />

chromium on TNF-α secretion in peripheral blood mononuclear cells (PBMCs)<br />

isolated from normal human blood. HG resulted in a significant increase in<br />

TNF-α secretion in PBMCs. Supplementation with chromium caused a


Trivalent Chromium Supplementation 305<br />

TNF-α Secretion (pg/ml)<br />

2700<br />

2500<br />

2300<br />

2100<br />

1900<br />

1700<br />

1500<br />

*<br />

#<br />

**<br />

FIGURE 20.1 Effect of high glucose and Cr 3+ on TNF-α secretion in PBMCs isolated<br />

from normal human volunteers. Values represent mean ± SE (n = 4). Differences in values<br />

(* versus**, ** versus ##) are significant (p


306 Oxidative Stress and Inflammatory Mechanisms<br />

HbA 1 (%)<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

#<br />

##<br />

*<br />

+ 50 mM Glucose<br />

FIGURE 20.3 Effects of different chromium concentrations on hemoglobin glycation in<br />

high glucose-treated erythrocytes. Values represent mean ± SE (n = 4). Differences in<br />

values (# versus ##, ## versus **, ## versus ***, and ## versus ****) are significant (p<br />


Trivalent Chromium Supplementation 307<br />

Lipid Peroxidation<br />

(nmol malonidialdehyde/ml cells)<br />

1.2<br />

1.1<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

*<br />

#<br />

*<br />

#<br />

FIGURE 20.4 Effect of chromium on lipid peroxidation in erythrocytes treated with<br />

different concentrations of glucose. Values represent mean ± SE (n = 4). Differences in<br />

values (* versus #) are significant (p


308 Oxidative Stress and Inflammatory Mechanisms<br />

FIGURE 20.5 Proposed model for role of trivalent chromium supplementation in prevention<br />

of oxidative stress and vascular disease in diabetes.<br />

levels of pro-inflammatory cytokines and thereby improve insulin sensitivity and<br />

glycemic control among the diabetic patient population. If so, chromium supplementation<br />

may be used as an adjuvant therapy for reduction of vascular inflammation<br />

and CVD in diabetes.<br />

ACKNOWLEDGMENTS<br />

The author is supported by a grant from NIDDK and the Office of Dietary<br />

Supplements of the National Institutes of Health (RO1 DK064797), and thanks<br />

Georgia Morgan for excellent editing of this manuscript.<br />

REFERENCES<br />

Diabetes<br />

Hyperglycemia Ketosis<br />

ROS<br />

(oxidative stress)<br />

Trivalent<br />

Chromium<br />

Protein Kinases<br />

MAPK Kinases (p38), NF-κB<br />

TNF-α, IL-6 Secretion<br />

Expression of Adhesion Molecules<br />

Monocyte-endothelial Cell Adhesion<br />

Vascular Inflammation<br />

Cardiovascular Disease<br />

1. Klein R. Hyperglycemia and microvascular and macrovascular disease in diabetes.<br />

Diab Care 18, 258, 1995.<br />

2. Rader DJ. Inflammatory markers of coronary risk. New Engl J Med 343, 1179,<br />

2000.<br />

3. Ming WF, Bersani L, and Mantovani A. Tumor necrosis factor-α is chemotactic<br />

for monocytes and polymorphonuclear leukocytes. J Immunol 138, 1469, 1987.<br />

4. Jain SK, Kannan K, Lim G, McVie R, and Bocchini JA. Hyperketonemia increases<br />

TNF-α secretion in cultured U937 monocytes and type-1 diabetic patients. Diabetes<br />

51, 2287, 2002.


Trivalent Chromium Supplementation 309<br />

5. Lechleitner M, Koch T, Herold M, Dzien A, and Hoppichler F. Tumour necrosis<br />

factor-alpha plasma level in patients with type I diabetes mellitus and its association<br />

with glycaemic control and cardiovascular risk factors. J Intern Med 248,<br />

67, 2000.<br />

6. Jain SK et al. Elevated blood interleukin-6 levels in hyperketonemic type 1 diabetic<br />

patients and secretion by acetoacetate-treated cultured U937 monocytes. Diab<br />

Care 26, 2139, 2003.<br />

7. Hussain MF et al. Elevated serum levels of macrophage-derived cytokines precede<br />

and accompany the onset of IDDM. Diabetologia 39, 60, 1996.<br />

8. Cavallo MG et al. Cytokines in sera from insulin-dependent diabetic patients at<br />

diagnosis. Clin Exp Immunol 86, 256, 1991.<br />

9. Ohno Y, Aoki N, and Nishimura A. In vitro production of interleukin-1, interleukin-<br />

6, and tumor necrosis factor-α in insulin-dependent diabetes mellitus. J Clin<br />

Endocrinol Metab 77, 1072, 1993.<br />

10. Morohoshi M, Fujisawa K, Uchimura I, and Numano F. The effect of glucose and<br />

advanced glycosylation end products on IL-6 production by human monocytes.<br />

Ann NY Acad Sci 748, 562, 1995.<br />

11. Andreozzi, F et al. Plasma interleukin-6 levels are independently associated with<br />

insulin secretion in a cohort of Italian Caucasian nondiabetic subjects. 55, 2021,<br />

2006.<br />

12. Kern PA, Ranganathan S, Li C, Wood L, and Ranganathan G. Adipose tissue<br />

tumor necrosis factor and interleukin-6 expression in human obesity and insulin<br />

resistance. Am J Physiol 280, E745, 2001.<br />

13. Halse R, Pearson SL, McCormack JG, Yeaman SJ, and Taylor R. Effects of tumor<br />

necrosis factor-α on insulin action in cultured human muscle cells. Diabetes 50,<br />

1102, 2001.<br />

14. Bruun JM, Verdich C, Toubro S, Astrup A, and Richelsen B. Association between<br />

measures of insulin sensitivity and circulating levels of interleukin-8, interleukin-<br />

6 and tumor necrosis factor-alpha: effect of weight loss in obese men. Eur J<br />

Endocrinol. 148, 535, 2003.<br />

15. Vozarova B et al. The interleukin-6 (–174) G/C promoter polymorphism is associated<br />

with type 2 diabetes mellitus in Native Americans and Caucasians. Hum<br />

Genet 112, 409, 2003.<br />

16. Rotter V, Nagaev I, and Smith U. Interleukin-6 induces insulin resistance in 3T3-<br />

L1 adipocytes and is like IL-8 and tumor necrosis factor-α overexpressed in human<br />

fat cells from insulin-resistant subjects. J Biol Chem 46, 45777, 2003.<br />

17. Gamble JR, Harlan JM, Klebanoff SF, and Vadas MA. Stimulation of the adherence<br />

of neutrophils to umbilical vein and endothelium by human recombinant<br />

tumor necrosis factor. Proc Natl Acad Sci USA 82, 8667, 1985.<br />

18. Cavender D, Saegusa Y, and Ziff M. Stimulation of endothelial cell binding of<br />

lymphocytes by tumor necrosis factor. J Immunol 139, 1855, 1987.<br />

19. Huber SA, Sakkinen P, Conze D, Hardin N, and Tracy R. Interleukin-6 exacerbates<br />

early atherosclerosis in mice. Arterioscler Thromb Vasc Biol 19, 2364, 1999.<br />

20. Zhang WJ and Frei B. Intracellular metal ion chelators inhibit TNF-alpha-induced<br />

SP-1 activation and adhesion molecule expression in human aortic endothelial<br />

cells. Free Radic Biol Med 34, 674, 2003.<br />

21. Cardelli M, Andreozzi F, Laretta E et al. Plasma interleukin-6 levels are increased<br />

in subjects with impaired glucose tolerance but not in those with impaired fasting<br />

glucose in a cohort of Italian Caucasians. Diab Metab Res Rev 23, 141, 2007.


310 Oxidative Stress and Inflammatory Mechanisms<br />

22. Granger DN, Vowinked T, and Petnehazy T. Modulation of the inflammatory<br />

reponse in cardiovascular disease. Hypertension 43, 924, 2004.<br />

23. Kim JA, Berliner JA, Natarajan RD, and Nadler JL. Evidence that glucose<br />

increases monocyte binding to human aortic endothelial cells. Diabetes 43, 1103,<br />

1994.<br />

24. Carley W et al. Distinct ICAM-1 forms and expression pathways in synovial<br />

microvascular endothelial cells. Cell Mol Biol 45, 79, 1999.<br />

25. Natarajan R, Reddy MA, Malik KU, Fatima S, and Khan BV. Signaling mechanisms<br />

of nuclear factor-kappa-b-mediated activation of inflammatory genes by<br />

13-hydroperoxyoctadecadienoic acid in cultured vascular smooth muscle cells.<br />

Arterioscler Thromb Vasc Biol 21, 1408, 2001.<br />

26. Bharti AC and Aggrawal BB. Nuclear factor-kappa B and cancer: its role in the<br />

prevention and therapy. Biochem Pharmacol 64, 883, 2002.<br />

27. Li N and Karin M. Is NF-kB the sensor of oxidative stress? FASEB J 13, 1137,<br />

1999.<br />

28. Suzuki YJ, Forman HJ, and Sevanian A. Oxidants as stimulators of signal transduction.<br />

Free Rad Biol Med 22, 269, 1997.<br />

29. Brownlee M. The pathogenesis of diabetic complications: a unifying mechanism.<br />

Diabetes 54, 1615, 2005.<br />

30. Igarashi M et al. Glucose or diabetes activates p38 mitogen-activated protein<br />

kinase via different pathways. J Clin Invest 103, 185, 1999.<br />

31. Guha M, Bai W, Nadler JL, and Natarajan R. Molecular mechanisms of tumor<br />

necrosis factor α gene expression in monocytic cells via hyperglycemia-induced<br />

oxidant stress-dependent and independent pathways. J Biol Chem 275, 17728,<br />

2000.<br />

32. Reddy MA et al. The oxidized lipid and lipoxygenase product 12(S)-hydroxyeicosatetraenoic<br />

acid induces hypertrophy and fibronectin transcription in vascular<br />

smooth muscle cells via p38 MAPK and cAMP response element-binding protein<br />

activation: mediation of angiotensin II effects. J Biol Chem 277, 9920, 2002.<br />

33. Daniel PB, Walker WH, and Habener JF. Cyclic AMP signaling and gene regulation.<br />

Ann Rev Nutr 18, 353, 1988.<br />

34. Chen CH, Zhang DH, LaPorte JM, and Ray A. Cyclic AMP activates p38 mitogenactivated<br />

protein kinase in Th2 cells, phosphorylation of GATA-3 and stimulation<br />

of Th2 cytokine gene expression. J Immunol 165, 5597, 2000.<br />

35. Serkkola E and Hurme M. Activation of NF-kappa B by cAMP in human myeloid<br />

cells. FEBS Lett 234, 327, 1993.<br />

36. Ollivier V, Parry GC, Cobb RR, de Prost D, and Mackman N. Elevated cyclic<br />

AMP inhibits NF-kappa-B-mediated transcription in human monocytic cells and<br />

endothelial cells. J Biol Chem 271, 20828, 1996.<br />

37. Treebak JT et al. AMP mediated AS-160 phosphorylation in skeletal muscle is<br />

dependent on AMPK catalytic and regulatory subunits. Diabetes 55, 2051, 2006.<br />

38. Singh U, Devraj S, and Jialal, I. Vitamin E, oxidative stress and inflammation.<br />

Ann Rev Nutr 25, 151, 2005.<br />

39. Vincent JB. Recent advances in the nutritional biochemistry of trivalent chromium.<br />

Proc Nutr Soc 63, 41, 2004.<br />

40. Vincent JB. Mechanisms of chromium action: low-molecular weight chromiumbinding<br />

substance. J Am Coll Nutr 18, 6, 1999.<br />

41. Offenbacher EG and Pi-Sunyer FX. Beneficial effect of chromium-rich yeast on<br />

glucose tolerance and blood lipids in elderly subjects. Diabetes 29, 919, 1980.


Trivalent Chromium Supplementation 311<br />

42. Cefalu WT and Hu FB. Role of chromium in human health and in diabetes. Diab<br />

Care 27, 2741, 2004<br />

43. Anderson RA. Chromium, glucose intolerance and diabetes. J Am Coll Nutr 17,<br />

548, 1998.<br />

44. Chromium and Diabetes Workshop, Office of Dietary Supplements, National<br />

Institutes of Health, Bethesda, 1999. http://ods.nih.gov/news/conferences/chromium_<br />

diabetes. html<br />

45. Jeejeebhoy KN. Chromium and parenteral nutrition. J Trace Elem Exp Med 12,<br />

85, 1999.<br />

46. Morris BW et al. Chromium homeostasis in patients with type II (NIDDM)<br />

diabetes. J Trace Elem Med Biol 13, 57, 1999.<br />

47. Department of Foods and Nutrition, American Medical Association. Guidelines<br />

for essential trace element preparations for parenteral use: statement by an expert<br />

panel. JAMA 241, 2051, 1979.<br />

48. Abraham AS et al. The effect of chromium on established atherosclerotic plaques<br />

in rabbits. Am J Clin Nutr 33, 2294, 1980.<br />

49. Abraham AS, Brooks BA, and Eylath U. Chromium- and cholesterol-induced<br />

atherosclerosis in rabbits. Ann Nutr Metab 35, 203, 1991.<br />

50. Stoecker BJ et al., Eds. Modern Nutrition in Health and Disease, Williams &<br />

Wilkins, Baltimore, 1999, p. 277.<br />

51. Panel on Micronutrients, Food and Nutrition Board, National Academy of Sciences.<br />

Dietary reference intakes for vitamin A, vitamin K, arsenic, boron, chromium,<br />

copper, iodine, iron, manganese, molybdenum, nickel, silicon, vanadium,<br />

and zinc. National Academy Press, Washington, 2001.<br />

52. Anderson RA, Bryden NA, and Polansky MM. Dietary chromium intake, freely<br />

chosen diets, institutional diet, and individual foods. Biol Trace Elem Res 32, 117,<br />

1992.<br />

53. Kozlovsky AS et al. Effects of diets high in simple sugars on urinary chromium<br />

losses. Metabolism 35, 515, 1986.<br />

54. Anderson RA and Kozlovsky AS. Chromium intake, absorption and excretion of<br />

subjects consuming self-selected diets. Am J Clin Nutr 41, 1177, 1985.<br />

55. Anderson RA. Chromium, in Trace Elements in Human and Animal Nutrition,<br />

Mertz W, Ed, Academic Press, Orlando, 1987, p. 225.<br />

56. Rajpathak S et al. Lower toenail chromium in men with diabetes and cardiovascular<br />

disease compared with healthy men. Diab Care 27, 2211, 2004.<br />

57. Pineau A, Guillard O, and Risse JF. A study of chromium in human cataractous<br />

lenses and whole blood of diabetics, senile, and normal population. Biol Trace<br />

Elem Res 32, 1338, 1992.<br />

58. Anderson RA et al. Elevated intakes of supplemental chromium improve glucose<br />

and insulin variables in individuals with type 2 diabetes. Diabetes 46, 1786, 1997.<br />

59. Mossop RT. Effects of chromium III on fasting blood glucose, cholesterol and<br />

cholesterol HDL levels in diabetics. Cent Afr J Med 29, 80, 1983.<br />

60. Cheng N et al. Follow-up survey of people in China with type 2 diabetes mellitus<br />

consuming supplemental chromium. J Trace Elem Exp Med 12, 55, 1999.<br />

61. Jovanovic-Peterson L, Gutierrez M., and Peterson, C. Chromium supplementation<br />

for gestational diabetes women improves glucose tolerance and decreases hyperinsulinemia.<br />

Diab Care 45 (Suppl. 2), 237A, 1996.


312 Oxidative Stress and Inflammatory Mechanisms<br />

62. Ravina, A, Slezak, L, Mirsky, N, Bryden, NA, and Anderson, RA. Reversal of<br />

corticosteroid-induced diabetes mellitus with supplemental chromium. Diab Med<br />

16, 164, 1999.<br />

63. Abraham AS, Brooks BA, and Eylat U. The effect of chromium supplementation<br />

on serum glucose and lipids in patients with non-insulin dependent diabetes<br />

mellitus. Med Clin Exp 41, 768, 1992.<br />

64. Ravina A et al. Clinical use of the trace element chromium (III) in the treatment<br />

of diabetes mellitus. J Trace Elem Exp Med 8, 183, 1995.<br />

65. Fox GN and Sabovic Z. Chromium picolinate supplementation for diabetes mellitus.<br />

J Fam Pract 46, 83, 1998.<br />

66. Anderson RA. Chromium and parenteral nutrition. Nutrition 11 (Suppl 1), 83,<br />

1995.<br />

67. Lee NA and Reasner CA. Beneficial effect of chromium supplementation on serum<br />

triglyceride levels in NIDDM. Diab Care 17, 1449, 1994.<br />

68. Lamson DM and Plaza SM. The safety and efficacy of high-dose chromium.<br />

Alternative Med Rev 7, 218, 2002.<br />

69. Guallar EJF et al. Toenail chromium and risk of myocardial infarction. Am J<br />

Epidemiol 162, 157, 2005.<br />

70. Jain SK. Hyperglycemia can cause membrane lipid peroxidation and osmotic<br />

fragility in human red blood cells. J Biol Chem 264, 21340, 1989.<br />

71. Rajeswari P, Natarajan R, Nadler JL, and Kumar, D. Glucose induces lipid peroxidation<br />

and inactivation of membrane associated iron transport enzymes in<br />

human erythrocytes in vivo and in vitro. J. Cell Physiol 149, 100, 1991.<br />

72. Natarajan R, Lanting L, Gonzales N, and Nadler J. Formation of an F2-isoprostane<br />

in vascular smooth muscle cells by elevated glucose and growth factors. Am J<br />

Physiol 271, 159, 1996.<br />

73. Giugliano D, Paolisso G, and Ceriello A. Oxidative stress and diabetic vascular<br />

complications. Diab Care 19, 257, 1996.<br />

74. Jain SK, McVie R, Duett J, and Herbst JJ. Erythrocyte membrane lipid peroxidation<br />

and glycosylated hemoglobin in diabetes. Diabetes 38, 1539, 1989.<br />

75. Hunter SJ et al. Demonstration of glycated insulin in human diabetic plasma and<br />

decreased biological activity assessed by euglycemic-hyperinsulinemic clamp<br />

technique in humans. Diabetes 52, 492, 2003.<br />

76. Jain SK and Palmer M. The effect of oxygen radical metabolites and vitamin E<br />

on glycosylation of proteins. Free Rad Biol Med 22, 593, 1997.<br />

77. Jain SK and Kannan K. Chromium chloride inhibits oxidative stress and TNF-α<br />

secretion caused by exposure to high glucose in cultured monocytes. Biochem<br />

Biophys Res Commun 289, 687, 2001.<br />

78. Bahijiri SM, Mira SA, Mufti AM, and Ajabnoor MA. The effects of inorganic<br />

chromium and brewer's yeast supplementation on glucose tolerance, serum lipids<br />

and drug dosage in individuals with type 2 diabetes. Saudi Med J 21, 831, 2000.<br />

79. Vinson JA, Mandarano MA, Shuta DL, Bagchi M, and Bagchi D. Beneficial effects<br />

of a novel IH636 grape seed proanthocyanidin extract and a niacin-bound chromium<br />

in a hamster atherosclerosis model. Mol Cell Biochem 240, 99, 2002.<br />

80. Preuss HG, Montamarry S, Echard B, Scheckenbach R, and Bagchi D. Long-term<br />

effects of chromium, grape seed extract, and zinc on various metabolic parameters<br />

of rats. Mol Cell Biochem 223, 95, 2001.


Trivalent Chromium Supplementation 313<br />

81. Cheng HH, Lai MH, Hou WC, and Huang CL. Antioxidant effects of chromium<br />

supplementation with type 2 diabetes mellitus and euglycemic subjects. J Agric<br />

Food Chem 52, 1385, 2004.<br />

82. Anderson RA et al. Potential antioxidant effects of zinc and chromium supplementation<br />

in people with type 2 diabetes mellitus. J Am Coll Nutr 20, 212, 2001.<br />

83. Jain SK, McVie R, and Bocchini JA. Ketosis, oxidative stress and type 1 diabetes.<br />

Pathophysiology 13, 163, 2006.<br />

84. Jain SK and Lim G. Chromium chloride inhibits TNF-α and IL-6 secretion in<br />

isolated human blood monoclear cells exposed to high glucose. Horm Metab Res<br />

38, 60, 2006.<br />

85. Jain SK, Patel, P, and Rogier K. Trivalent chromium inhibits protein glycation<br />

and lipid peroxidation in high glucose-treated erythrocytes. Antioxidant Redox<br />

Signaling 8, 238, 2006.


Index<br />

A<br />

Abdominal obesity thresholds, 10<br />

gender comparisons, 10<br />

Adenosine triphosphate, uncoupling protein 2,<br />

pancreatic beta cell function<br />

beta cell mass, survival, 218–219<br />

insulin secretion, 216–218<br />

Adenosine triphosphate III<br />

initial report, metabolic syndrome<br />

definition, 5<br />

metabolic syndrome, 6<br />

revised criteria/guidelines, 7–9<br />

Adhesion molecules, clinical marker of<br />

inflammation, 152<br />

Adipokines, 116–118<br />

Adiponectin<br />

clinical marker of inflammation, 147–148<br />

oligomeric composition, 167–176<br />

adipocytes, 168–169<br />

central nervous system effects,<br />

172–173<br />

multimers, biological activities of,<br />

169–171<br />

thermogenesis, 172<br />

weight loss results, 171–172<br />

Adipose tissue, nutritional inflammation<br />

modulation, 231–232<br />

Alpha lipoic acid, 273–288. See also<br />

Lipoic acid<br />

adipocyte differentiation regulation, 280<br />

MAPK signaling pathways, 281–283<br />

transcription factor modulation,<br />

281–282<br />

diabetes mellitus prevention, 261–272<br />

hypothalamic AMPK, 277–279<br />

insulin resistance, 289–300<br />

etiology of insulin resistance, 292<br />

nutriceutical interventions, 292–295<br />

skeletal muscle glucose transport<br />

system, regulation, 291<br />

structure, 274<br />

American Association of Clinical<br />

Endocrinologists<br />

metabolic syndrome guidelines, 5–7<br />

Anti-inflammatory effect, insulin, proinflammatory<br />

effect,<br />

macronutrients, balance,<br />

metabolic syndrome, 15–32<br />

Antioxidants<br />

oxidative stress, 71–92<br />

childbirth, 75–78<br />

perinatal asphyxia, 76–77<br />

resuscitation, use of pure oxygen,<br />

77–78<br />

human milk, 82–84<br />

in overweight, obesity, 38–39<br />

perinatal period, 72–73<br />

pregnancy, 74–75<br />

fetal development, reactive oxygen<br />

species, 74<br />

preeclampsia, 75<br />

premature infants, 78–82<br />

bronchopulmonary dysplasia, 79<br />

neonatal necrotizing enterocolitis,<br />

80<br />

periventricular leukomalacia, 80–81<br />

retinopathy of prematurity, 81–82<br />

reactive oxygen species, tissueproduced,<br />

72<br />

perinatal period, oxidative stress, 71–92<br />

Atherosclerosis, inflammation, 139–166<br />

adhesion molecules, 152<br />

adiponectin, 147–148<br />

C-reactive protein, 144–146, 157–159<br />

cardiovascular risk, 141–143<br />

clinical markers, 143–152<br />

hemostatic factors, 159–160<br />

hemostatic parameters, 150–151<br />

interleukin-1, 152<br />

interleukin-6, 152<br />

interleukin-10, 152<br />

leptin, 146–147<br />

modification, 156–160<br />

resistin, 151–152<br />

tumor necrosis factor-alpha, 148–150, 159<br />

visfatin, 152<br />

Atherosclerotic plaque destabilization, lipidinduced<br />

macrophage death,<br />

251–260<br />

atherosclerosis, 254–255<br />

315


316 Oxidative Stress and Inflammatory Mechanisms<br />

eicosanoids, 258<br />

programmed cell death, 253–254<br />

therapies, 258<br />

triacylglycerols, 255–257<br />

ATP. See Adenosine triphosphate<br />

B<br />

Beta cell failure, 115–116<br />

Body weight, obesity, 33–46<br />

abdominal thresholds, 10<br />

adipocyte differentiation, 273–288<br />

alpha lipoic acid, 261–272<br />

antioxidants, 38–39<br />

atherosclerosis, 139–166<br />

cardiovascular risk, 141–143<br />

diabetes mellitus, 107–122, 261–272<br />

early life nutritional experience, correlation,<br />

62–63<br />

eicosapentaenoic acid, 41<br />

endothelial dysfunction, 261–272<br />

F 2-isoprostanes, 34–38<br />

fatty cell biology, 274–277<br />

hypothalamic AMPK suppression, 277–279<br />

increasing rates of, 108–110<br />

inflammation, 20–22, 94–96, 139–166<br />

insulin, macronutrient balance, 20–22<br />

insulin resistance, 94–96<br />

International Diabetes Federation,<br />

abdominal threshold criteria, 10<br />

linoleic acid, 289–300<br />

lipoic acid, 273–300<br />

macronutrient, insulin balance, 20–22<br />

maternal, 93–106<br />

metabolic programming, 49–62<br />

nationality/gender comparisons, abdominal<br />

thresholds, 10<br />

nutrigenomics, 107–122<br />

oligomeric composition, adiponectin,<br />

167–176<br />

omega-3 polyunsaturated fatty acids, 39–41<br />

oxidative stress, 33–46<br />

reduction, therapeutic targets, 38–41<br />

therapeutic targets for reduction, 38–41<br />

Bronchopulmonary dysplasia, 79<br />

C<br />

C-reactive protein, clinical marker of<br />

inflammation, 144–146, 157–159<br />

Cardiovascular disease, 111–113<br />

diabetes mellitus, post-prandial endothelial<br />

dysfunction<br />

fasting hyperglycemia, 124–125<br />

hyperglycemic spikes, 124–127<br />

mechanisms, 127–128<br />

oxidative, nitrosative stress, 128–131<br />

post-prandial hyperglycemia,<br />

128–131<br />

post-prandial hyperglycemia, 124–127<br />

nutritional inflammation modulation,<br />

228–229<br />

Central nervous system effects, adiponectin,<br />

172–173<br />

Childbirth, oxidative stress, 75–78<br />

perinatal asphyxia, 76–77<br />

resuscitation, use of pure oxygen, 77–78<br />

Chinese national origin, abdominal obesity<br />

thresholds, 10<br />

Chromium supplementation, 301–312<br />

diabetes mellitus, 303–307<br />

oxidative stress, 303<br />

pro-inflammatory cytokines<br />

diabetes mellitus, 302–303<br />

vascular inflammation, 302–303<br />

vascular inflammation, 303<br />

CLA. See Conjugated linoleic acid<br />

Clinical markers, inflammation, 143–152,<br />

156–160<br />

adhesion molecules, 152<br />

adiponectin, 147–148<br />

C-reactive protein, 144–146, 157–159<br />

hemostatic parameters, 150–151<br />

interleukin-1, 152<br />

interleukin-6, 152<br />

interleukin-10, 152<br />

leptin, 146–147<br />

resistin, 151–152<br />

tumor necrosis factor-alpha, 148–150, 159<br />

visfatin, 152<br />

Conjugated linoleic acid, insulin resistance,<br />

289–300<br />

etiology of insulin resistance, 292<br />

nutriceutical interventions, 292–295<br />

skeletal muscle glucose transport system,<br />

regulation, 291<br />

D<br />

Definition of metabolic syndrome, 3–14<br />

abdominal obesity thresholds, 10<br />

adenosine triphosphate III<br />

initial report, 5<br />

metabolic syndrome, 6


Index 317<br />

revised criteria/guidelines, 7–9<br />

American Association of Clinical<br />

Endocrinologists, 5–7<br />

European Group for Study of Insulin<br />

Resistance Syndrome, 7–8, 12<br />

features of metabolic syndrome definitions,<br />

12<br />

INTERHEART project, 3, 9<br />

International Diabetes Federation<br />

definition, 8–9<br />

metabolic syndrome criteria, 10<br />

metabolic syndrome definition, 12<br />

nationality/gender comparisons, abdominal<br />

obesity thresholds, 10<br />

rationale for defining, 3–5<br />

World Health Organization<br />

metabolic syndrome definition, 12<br />

metabolic syndrome diagnostic criteria,<br />

6<br />

World Health Organization definition, 5<br />

Diabetes mellitus<br />

alpha lipoic acid, 261–272<br />

nutrigenomics, metabolic syndrome,<br />

107–122<br />

adipokines, 116–118<br />

beta cell failure, 115–116<br />

cardiometabolic risk, 111–113<br />

criteria for diagnosing metabolic<br />

syndrome, 113<br />

patient assessment, 109<br />

pre-diabetes mellitus, 111–113<br />

nutritional inflammation modulation, 229<br />

post-prandial endothelial dysfunction,<br />

123–138<br />

cardiovascular disease<br />

fasting hyperglycemia, 124–125<br />

hyperglycemic spikes, 124–127<br />

mechanisms, 127–128<br />

oxidative, nitrosative stress,<br />

128–131<br />

post-prandial hyperglycemia,<br />

128–131<br />

post-prandial hyperglycemia,<br />

124–127<br />

trivalent chromium supplementation,<br />

302–307<br />

vascular remodeling, 195–210<br />

Akt/PKB signaling pathway, 202<br />

dual PPAR activators, 204<br />

e-Jun N-terminal kinase signaling, 201<br />

extracellular signal-regulated kinase 1<br />

signaling, 199–200<br />

inflammation, vascular, 203–204<br />

MAPK signaling, 198<br />

p38 signaling, 200–201<br />

PI3K signaling pathway, 201–202<br />

PPAR-gamma, 199–200<br />

reactive oxygen species, 202–203<br />

Dietary fatty acids, metabolic syndrome,<br />

inflammation, 243–250<br />

Dihydrolipoic acid, structure, 262, 274<br />

Dysplasia, bronchopulmonary, 79<br />

E<br />

Early life nutritional modifications, metabolic<br />

syndrome, 47–70<br />

metabolic programming, 49–62<br />

generational effects, 59–62<br />

high carbohydrate model, 54–58<br />

immediate effects, 56–58<br />

immediate post-natal period, 53–54<br />

persistent effects in adulthood, 58–59<br />

pre-natal metabolic programming,<br />

50–53<br />

EGIR syndrome. See European Group for Study<br />

of Insulin Resistance Syndrome<br />

Eicosapentaenoic acid, oxidative stress, in<br />

overweight, obesity, 41<br />

Energy expenditure<br />

alpha lipoic acid, 273–288<br />

adipocyte differentiation regulation, 280<br />

MAPK signaling pathways,<br />

281–283<br />

transcription factor modulation,<br />

281–282<br />

hypothalamic AMPK, 277–279<br />

Enterocolitis, neonatal, necrotizing, 80<br />

EPA. See Eicosapentaenoic acid<br />

European Group for Study of Insulin Resistance<br />

Syndrome, 7–8, 12<br />

European national origin, abdominal obesity<br />

thresholds, 10<br />

F<br />

F 2-isoprostanes, oxidative stress in overweight,<br />

obesity, 34–38<br />

Fasting hyperglycemia, cardiovascular disease,<br />

124–125<br />

Fatty acids, metabolic syndrome, inflammation,<br />

243–250<br />

Features of metabolic syndrome definitions, 12


318 Oxidative Stress and Inflammatory Mechanisms<br />

G<br />

Gender comparisons, abdominal obesity<br />

thresholds, 10<br />

Generational effects, early life nutritional<br />

modifications, 59–62<br />

Glucose intolerance, maternal obesity,<br />

pregnancy, inflammation, 93–106<br />

fetal growth, 100<br />

interventions, 101–103<br />

H<br />

High carbohydrate model, early life nutritional<br />

modifications, 54–58<br />

Human islets, UCP2 protein expression, 213<br />

Hyperglycemic spikes, cardiovascular disease,<br />

124–127<br />

Hypertension, vascular remodeling, 195–210<br />

Akt/PKB signaling pathway, 202<br />

dual PPAR activators, 204<br />

e-Jun N-terminal kinase signaling, 201<br />

extracellular signal-regulated kinase 1<br />

signaling, 199–200<br />

extracellular signal-regulated kinase 2<br />

signaling, 199–200<br />

inflammation, vascular, 203–204<br />

MAPK signaling, 198<br />

p38 signaling, 200–201<br />

PI3K signaling pathway, 201–202<br />

PPAR-gamma, 199–200<br />

reactive oxygen species, 202–203<br />

Hypocaloric diets, inflammation modulation,<br />

232–237<br />

I<br />

IDF. See International Diabetes Federation<br />

Inflammation<br />

in atherosclerosis, 139–166<br />

dietary fatty acids, insulin signaling,<br />

244–246<br />

insulin, anti-inflammatory effect, 15–32<br />

metabolic syndrome, 15–32, 227–242<br />

in pregnancy, 93–106<br />

pro-inflammatory effect, macronutrients,<br />

15–32<br />

trivalent chromium supplementation,<br />

301–312<br />

in type 2 diabetes, 123–138<br />

vascular, 203–204, 301–312<br />

Insulin resistance, 16, 289–300<br />

etiology of insulin resistance, 292<br />

inflammatory hypothesis, 22<br />

macronutrient, metabolic effects, 16–17<br />

metabolic effects, 16–17<br />

nutriceutical interventions, 292–295<br />

skeletal muscle glucose transport system,<br />

regulation, 291<br />

Insulin resistance syndrome, usage of term, 7<br />

Insulin signaling, dietary fatty acids, 243–250<br />

Insulin-stimulated reactive oxygen species,<br />

signal transduction, 177–194<br />

cellular insulin stimulation, H 2O 2<br />

generation, 181<br />

cellular tyrosine kinase signaling, reactive<br />

oxygen species as second<br />

messengers, 179<br />

insulin-stimulated H 2O 2 generation, proteintyrosine<br />

phosphatase, 1B<br />

regulation, 182<br />

novel targets, 186<br />

Nox4, signaling regulation, insulin action<br />

cascade, 184–185<br />

protein-tyrosine phosphatase, 180–181<br />

reversible tyrosine phosphorylation,<br />

178–179<br />

INTERHEART project, 3, 9<br />

Interleukin-1, clinical marker of inflammation,<br />

152<br />

Interleukin-6, clinical marker of inflammation,<br />

152<br />

Interleukin-10, clinical marker of inflammation,<br />

152<br />

International Diabetes Federation<br />

metabolic syndrome, 12<br />

criteria, 10<br />

definition, 8–9<br />

International Diabetes Federation criteria,<br />

thresholds for abdominal obesity,<br />

10<br />

J<br />

Japanese national origin, abdominal obesity<br />

thresholds, 10<br />

L<br />

Leptin, clinical marker of inflammation,<br />

146–147<br />

Leukomalacia, periventricular, 80–81


Index 319<br />

Lipid-induced macrophage death,<br />

atherosclerotic plaque<br />

destabilization, 251–260<br />

atherosclerosis, 254–255<br />

eicosanoids, 258<br />

programmed cell death, 253–254<br />

therapies, 258<br />

triacylglycerols, 255–257<br />

Lipoic acid, 273–288. See also alpha lipoic acid<br />

adipocyte differentiation regulation, 280<br />

MAPK signaling pathways, 281–283<br />

transcription factor modulation,<br />

281–282<br />

hypothalamic AMPK, 277–279<br />

insulin resistance, 289–300<br />

etiology of insulin resistance, 292<br />

nutriceutical interventions, 292–295<br />

skeletal muscle glucose transport<br />

system, regulation, 291<br />

structure, 262<br />

M<br />

Macronutrients, insulin balance, metabolic<br />

syndrome, 15–32<br />

inflammation, 25–26<br />

insulin resistance, 16<br />

inflammatory hypothesis, 22<br />

metabolic effects, 16–17<br />

non-metabolic actions of insulin, 17–20<br />

obesity, 20–22<br />

origin of inflammation, 22–25<br />

pathogenesis, 24<br />

MAPK signaling pathways, responses, 198<br />

Maternal obesity, 93–106<br />

glucose intolerance, inflammation, 93–106<br />

fetal growth, 100<br />

interventions, 101–103<br />

Metabolic syndrome, 107–122. See also<br />

diabetes mellitus<br />

criteria for diagnosing, 113<br />

definition of, 3–14 .See also under<br />

Definition of metabolic syndrome<br />

dietary fatty acids, 243–250 .See also under<br />

Dietary fatty acids<br />

inflammation modulation, 227–242 .See<br />

also under Inflammation<br />

macronutrients, insulin balance, 15–32 .See<br />

also under Macronutrients, insulin<br />

balance, metabolic syndrome<br />

nutrigenomics, diabetes mellitus, 107–122<br />

adipokines, 116–118<br />

beta cell failure, 115–116<br />

cardiometabolic risk, 111–113<br />

criteria for diagnosing metabolic<br />

syndrome, 113<br />

patient assessment, 109<br />

pre-diabetes mellitus, 111–113<br />

nutritional modifications, early life, 47–70.<br />

See also Nutritional modifications,<br />

early life<br />

Mitogen-activated protein kinase. See MAPK<br />

N<br />

Nationality/gender comparisons, abdominal<br />

obesity thresholds, 10<br />

Neonatal necrotizing enterocolitis, 80<br />

Non-metabolic actions of insulin, 17–20<br />

Nutrigenomics, diabetes mellitus, metabolic<br />

syndrome, 107–122<br />

adipokines, 116–118<br />

beta cell failure, 115–116<br />

cardiometabolic risk, 111–113<br />

criteria for diagnosing metabolic syndrome,<br />

113<br />

patient assessment, 109<br />

pre-diabetes mellitus, 111–113<br />

Nutritional modifications, early life, metabolic<br />

syndrome, 47–70<br />

metabolic programming, 49–62<br />

generational effects, 59–62<br />

high carbohydrate model, 54–58<br />

immediate effects, 56–58<br />

immediate post-natal period, 53–54<br />

persistent effects in adulthood, 58–59<br />

pre-natal metabolic programming,<br />

50–53<br />

O<br />

Obesity, 20–22, 33–46<br />

abdominal thresholds, 10<br />

adipocyte differentiation, 273–288<br />

alpha lipoic acid, 261–272<br />

antioxidants, 38–39<br />

atherosclerosis, 139–166<br />

adhesion molecules, 152<br />

adiponectin, 147–148<br />

C-reactive protein, 144–146, 157–159<br />

cardiovascular risk, 141–143<br />

clinical markers, inflammation, 143–152<br />

hemostatic factors, inflammation<br />

modification, 159–160<br />

hemostatic parameters, 150–151


320 Oxidative Stress and Inflammatory Mechanisms<br />

interleukin-1, 152<br />

interleukin-6, 152<br />

interleukin-10, 152<br />

leptin, 146–147<br />

modification of inflammation, 156–160<br />

resistin, 151–152<br />

tumor necrosis factor-alpha, 148–150,<br />

159<br />

visfatin, 152<br />

cardiovascular risk, 141–143<br />

diabetes mellitus, 107–122, 261–272<br />

early life nutritional experience, correlation,<br />

62–63<br />

early life nutritional modifications, 47–70<br />

eicosapentaenoic acid, 41<br />

endothelial dysfunction, 261–272<br />

F 2-isoprostanes, 34–38<br />

fatty cell biology, 274–277<br />

hypothalamic AMPK suppression, 277–279<br />

increasing rates of, 108–110<br />

inflammation, 20–22, 94–96, 139–166<br />

insulin, macronutrient balance, 20–22<br />

insulin resistance, 94–96<br />

International Diabetes Federation,<br />

abdominal threshold criteria, 10<br />

linoleic acid, 289–300<br />

lipoic acid, 273–300<br />

macronutrient, insulin balance, 20–22<br />

maternal, 93–106<br />

metabolic programming, 49–62<br />

nationality/gender comparisons, abdominal<br />

thresholds, 10<br />

nutrigenomics, 107–122<br />

oligomeric composition, adiponectin,<br />

167–176<br />

omega-3 polyunsaturated fatty acids, 39–41<br />

oxidative stress, 33–46<br />

reduction, therapeutic targets, 38–41<br />

therapeutic targets for reduction, 38–41<br />

Oligomeric composition, adiponectin, 167–176<br />

adipocytes, 168–169<br />

central nervous system effects, 172–173<br />

multimers, biological activities of, 169–171<br />

thermogenesis, 172<br />

weight loss results, 171–172<br />

Omega-3 polyunsaturated fatty acids, oxidative<br />

stress, in overweight, obesity,<br />

39–41<br />

Origin of inflammation, 22–25<br />

Oxidative stress<br />

antioxidants, 71–92<br />

childbirth, 75–78<br />

perinatal asphyxia, 76–77<br />

P<br />

resuscitation, use of pure oxygen,<br />

77–78<br />

human milk, 82–84<br />

perinatal period, 72–73<br />

pregnancy, 74–75<br />

fetal development, reactive oxygen<br />

species, 74<br />

preeclampsia, 75<br />

premature infants, 78–82<br />

bronchopulmonary dysplasia, 79<br />

neonatal necrotizing enterocolitis,<br />

80<br />

periventricular leukomalacia, 80–81<br />

retinopathy of prematurity, 81–82<br />

reactive oxygen species, tissueproduced,<br />

72<br />

diabetes mellitus, 123–138<br />

obesity, 33–46<br />

antioxidants, 38–39<br />

eicosapentaenoic acid, 41<br />

F 2-isoprostanes, 35–38<br />

as markers, 34–35<br />

omega-3 polyunsaturated fatty acids,<br />

39–41<br />

reduction, therapeutic targets, 38–41<br />

perinatal period, 71–92<br />

trivalent chromium supplementation, 303<br />

Pancreatic beta cell function, uncoupling<br />

protein 2, 211–224<br />

adenosine triphosphate<br />

beta cell mass, survival, 218–219<br />

insulin secretion, 216–218<br />

expression, regulatory factors of, 212–216<br />

reactive oxygen species production, beta cell<br />

mass, survival, 218–219<br />

Perinatal asphyxia, 76–77<br />

Perinatal period, oxidative stress, 71–92<br />

Periventricular leukomalacia, 80–81<br />

Peroxisome proliferator-activated receptors,<br />

hypertension, 195–210<br />

Post-natal period metabolic programming,<br />

53–54<br />

Post-prandial endothelial dysfunction, diabetes<br />

mellitus, 123–138<br />

cardiovascular disease<br />

fasting hyperglycemia, 124–125<br />

hyperglycemic spikes, 124–127<br />

mechanisms, 127–128<br />

oxidative, nitrosative stress, 128–131


Index 321<br />

post-prandial hyperglycemia,<br />

128–131<br />

post-prandial hyperglycemia, 124–127<br />

PPARs. See Peroxisome proliferator-activated<br />

receptors<br />

Pre-natal metabolic programming, 50–53<br />

Preeclampsia, oxidative stress, 75<br />

Pregnancy, 93–106<br />

glucose intolerance, inflammation, 93–106<br />

fetal growth, 100<br />

interventions, 101–103<br />

maternal obesity, glucose intolerance,<br />

inflammation, 93–106<br />

fetal growth, 100<br />

interventions, 101–103<br />

oxidative stress, 74–75<br />

fetal development, reactive oxygen<br />

species, 74<br />

preeclampsia, 75<br />

Premature infants, oxidative stress, 78–82<br />

bronchopulmonary dysplasia, 79<br />

neonatal necrotizing enterocolitis, 80<br />

periventricular leukomalacia, 80–81<br />

retinopathy of prematurity, 81–82<br />

Pro-inflammatory cytokines<br />

monocytes, inflammation modulation,<br />

230–231<br />

trivalent chromium supplementation<br />

diabetes mellitus, 302–303<br />

vascular inflammation, 302–303<br />

vascular inflammation, trivalent chromium<br />

supplementation, 302–303<br />

Pro-inflammatory effect, macronutrients, antiinflammatory<br />

effect, insulin,<br />

balance, metabolic syndrome,<br />

15–32<br />

R<br />

Rationale for defining metabolic syndrome, 3–5<br />

Resistin, clinical marker of inflammation,<br />

151–152<br />

Resuscitation in perinatal asphyxia, use of pure<br />

oxygen, 77–78<br />

Retinopathy of prematurity, 81–82<br />

S<br />

Signal transduction, insulin-stimulated reactive<br />

oxygen species, 177–194<br />

cellular insulin stimulation, H 2O 2<br />

generation, 181<br />

cellular tyrosine kinase signaling, reactive<br />

oxygen species as second<br />

messengers, 179<br />

insulin-stimulated H 2O 2 generation, proteintyrosine<br />

phosphatase, 1B<br />

regulation, 182<br />

novel targets, 186<br />

Nox4, signaling regulation, insulin action<br />

cascade, 184–185<br />

protein-tyrosine phosphatase, 180–181<br />

inhibition, 183<br />

reversible tyrosine phosphorylation,<br />

178–179<br />

South Asian national origin, abdominal obesity<br />

thresholds, 10<br />

Stress, oxidative, in overweight, obesity, 33–46<br />

antioxidants, 38–39<br />

eicosapentaenoic acid, 41<br />

F 2-isoprostane formation, 35<br />

F 2-isoprostanes, 35–38<br />

as markers, 34–35<br />

omega-3 polyunsaturated fatty acids, 39–41<br />

reduction, therapeutic targets, 38–41<br />

T<br />

Thermogenesis, adiponectin, 172<br />

Trivalent chromium supplementation, 301–312<br />

diabetes mellitus, 303–307<br />

oxidative stress, 303<br />

pro-inflammatory cytokines<br />

diabetes mellitus, 302–303<br />

vascular inflammation, 302–303<br />

vascular inflammation, 303<br />

Tumor necrosis factor-alpha, clinical marker of<br />

inflammation, 148–150, 159<br />

U<br />

UCP2 protein expression, human islets, 213<br />

Uncoupling protein 2<br />

pancreatic beta cell function, 211–224<br />

adenosine triphosphate<br />

beta cell mass, survival, 218–219<br />

insulin secretion, 216–218<br />

expression, regulatory factors of,<br />

212–216<br />

reactive oxygen species production, beta<br />

cell mass, survival, 218–219


322 Oxidative Stress and Inflammatory Mechanisms<br />

V<br />

Vascular inflammation<br />

pro-inflammatory cytokines, trivalent<br />

chromium supplementation,<br />

302–303<br />

trivalent chromium supplementation,<br />

302–303<br />

Vascular remodeling, hypertension, diabetes<br />

mellitus, 195–210<br />

Akt/PKB signaling pathway, 202<br />

dual PPAR activators, 204<br />

e-Jun N-terminal kinase signaling, 201<br />

extracellular signal-regulated kinase 1<br />

signaling, 199–200<br />

extracellular signal-regulated kinase 2<br />

signaling, 199–200<br />

inflammation, vascular, 203–204<br />

MAPK signaling, 198<br />

p38 signaling, 200–201<br />

PI3K signaling pathway, 201–202<br />

PPAR-gamma, 199–200<br />

reactive oxygen species, 202–203<br />

Visfatin, clinical marker of inflammation, 152<br />

W<br />

WHO. See World Health Organization<br />

World Health Organization<br />

metabolic syndrome<br />

definition, 12<br />

diagnostic criteria, 6<br />

metabolic syndrome definition, 5

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!